Abstract
The aminoacyl-tRNA synthetases are prominently known for their classic function in the first step of protein synthesis, where they bear the responsibility of setting the genetic code. Each enzyme is exquisitely adapted to covalently link a single standard amino acid to its cognate set of tRNA isoacceptors. These ancient enzymes have evolved idiosyncratically to host alternate activities that go far beyond their aminoacylation role and impact a wide range of other metabolic pathways and cell signaling processes. The family of aminoacyl-tRNA synthetases have also been suggested as a remarkable scaffold to incorporate new domains that would drive evolution and the emergence of new organisms with more complex function. Because they are essential, the tRNA synthetases have served as pharmaceutical targets for drug and antibiotic development. The recent unfolding of novel important functions for this family of proteins offers new and promising pathways for therapeutic development to treat diverse human diseases.
The aminoacyl-tRNA synthetases (aaRSs) comprise an ancient family of enzymes that are responsible for the first step of protein synthesis. This diverse set of proteins is united by a common aminoacylation reaction, which attaches an amino acid to its cognate tRNA. Each aaRS is adapted to activate a single amino acid (aa) via an adenylate intermediate (aa-AMP) and then covalently link it to one of the ribose hydroxyls at the 3′ end of a set of tRNA isoacceptors.
| (I) |
| (II) |
Identity elements within each tRNA serve as molecular determinants that confer specificity of the aaRS for tRNA [1]. Accurate amino acid selection is challenging for about half of the aaRSs and is compensated by a hydrolytic editing domain to clear mistakes [2]. Thus, each aaRS has been finely tuned through a lengthy evolutionary period to ensure accurate protein synthesis.
The aaRSs have also been recruited for a myriad of other idiosyncratic activities that are unrelated to their primary role in protein synthesis [3, 4]. These include supporting RNA splicing, cell signaling, as well as transcriptional and translational regulation. Paralogs of aaRSs also exist, which have lost their aminoacylation function [5]. However, in many cases, the aaRS maintains dual roles in protein synthesis and its alternate activities. Some of these activities are unleashed by variations in splicing events, proteolytic clevage, or post-translational modifications [6]. Remarkably, the incorporation of novel domains into members of the aaRS family has been evolutionarily linked with increased complexity that yields new organisms.
aaRS Classifications and Distinctions
The family of aaRSs has been divided into two classes based on their chemical properties, architecture of their catalytic domains and consensus sequences [7-9] (Table 1). Each class can be further sub-divided into subclasses based on their unique mechanistic properties, anticodon-binding domain characteristics, as well as the organization of conserved structural motifs (Table 1) [2, 9-13]. With only one exception, lysyl-tRNA synthetase (LysRS) [14], the division of the class 1 and 2 aaRSs has been strictly conserved.
Table 1.
| Class I | Class II |
|---|---|
|
| |
| Ia | IIa |
| LeuRS* | SerRS |
| IleRS* | ThrRS* |
| ValRS* | ProRS* |
| MetRS* | HisRS |
| CysRS | GlyRS (Dimer) |
| ArgRS | |
| IIb | |
| Ib | AspRS |
| GlnRS | AsnRS |
| GluRS | LysRS-II* |
| LysRS-I | |
| IIc | |
| Ic | PheRS* |
| TyrRS | AlaRS* |
| TrpRS | GlyRS (Tetramer) |
| PylRS | |
| SepRS | |
These aaRSs have a hydrolytic editing pathway to clear misactivated and mischarged amino acids [2].
A Rossmann ATP binding fold defines the canonical core of class I aaRSs, which carries out the aminoacylation reaction [15]. The active site is marked by two consensus sequences: “KMSKS” (Lys-Met-Ser-Lys-Ser) [16-18] and “HIGH” (His-Ile-Gly-His) [19] (Figure 1). The KMSKS sequence functions to stabilize ATP during amino acid activation [20], in addition to stabilizing the 3′ end of tRNA for amino acid transfer [20-23]. This short KMSKS peptide is found in loops that bridge the β-strand S6 and the α-helix H11 [16]. During amino acid activation, the KMSKS sequence moves approximately 8 Å from an “opened” to a “closed” conformation in order to bring the KMSKS polypeptide in close proximity to stabilize the pyrophosphate moiety of ATP [24]. The N-terminal HIGH sequence also plays a role in amino acid activation [19], whereby the histidines bind to and stabilize the phosphate backbone of ATP [25]. The HIGH sequence is found in loops between the β-strand S2 and the α-helix H3 [16]. Its movement is synchronous with the “opened” and “closed” conformations of the KMSKS sequence [24].
Figure 1.
Conserved structure motifs and ATP binding in class I and II aaRSs. A. Class I LeuRS enzyme [Thermus thermophilus (T. thermophilus)] binds to an extended conformation of a sulphamoyl analogue of leucyl-adenylate (black) within the synthetic active site. The conserved signature sequences “HIGH” (HMGH) and “KMSKS” (MSKSK) are highlighted in blue and green respectively (PDB: 2V0C) [240]. While these signature sequences can vary as shown in this example, they are readily recognizable in a sequence alignment of all class I aaRSs. B. Class II glycyl-tRNA synthetase (GlyRS) binds ATP (black) in a bent conformation. The three conserved sequences, motifs 1, 2, and 3 of class II aaRSs are highlighted orange, purple, and blue respectively (PDB: 1B76) [247].
In contrast to the class I enzymes, which tend to be monomeric, class II aaRSs are almost exclusively oligomeric and are usually in dimeric or tetrameric form. The topology of the class II aaRS catalytic core is only found in a few non–aaRS proteins and is comprised of seven antiparallel β-strands flanked by α-helices [7, 8, 26]. Three conserved sequence motifs define the class II aaRSs (Figure 1). Motif 1 [gφxxφxxpφφ where φ represents a hydrophobic amino acid] is located at the interface of the dimer. It has been hypothesized to provide a communication mechanism between the active sites of the two subunits [27]. Motif 2 [fRxe-h/rxxxFxxx(d/e)] and motif 3 [gφgφgφ(d/e)Rφφφφφ] comprise part of the aminoacylation active site [7, 8, 28]. In the past decade, two new aaRSs that activate O-phosphoserine (SepRS) and pyrrolysine (PylRS) have been added to the set of class II aaRSs [29, 30] (Table 1).
Historically, LysRS had been associated with the class II aaRSs, but a few surprising cases have been found in archaea that represent the class I enzymes [14]. For example, there is not a class II LysRS in Methanococcus maripaludis LysRS [14]. Rather, a 62 kDa class I protein has been identified as a LysRS. In just one organism (Methanosarcineae) identified so far, both the class I and II LysRSs co-exist within a single organism. In this case, it is hypothesized that the two LysRSs collaborate to promote the aminoacylation of the rare amino acid pyrrolysine [31].
The two distinct cores of the aaRS classes confer different mechanisms to carry out the two-step aminoacylation reaction. Class I aaRSs bind ATP in an extended conformation [16, 17, 32], while class II aaRSs bind ATP in a bent conformation [21, 33] (Figure 1) due to a difference in active site structure of the aaRS. In addition, the class I aaRSs bind the tRNA stem via the minor groove side [17], thus orienting the 2′-hydroxyl group of the A76 ribose for attachment of amino acid [34]. In contrast, class II aaRSs bind the tRNA stem via the major groove side [28], resulting in accessibility of the 3′-hydroxyl group for charging [34].
Some aaRSs have acquired insertions or appendages to their catalytic canonical cores. In many cases, these inserted domains enhance enzyme specificity and fidelity for the aminoacylation reaction [5]. About half of the aaRSs rely on an extra domain with a hydrolytic active site to edit mistakes that were made in the aminoacylation active site of the canonical core domain [2]. The connective polypeptide 1 (CP1) [35] hydrolytic editing domain ensures amino acid fidelity of the class I leucyl- (LeuRS), isoleucyl- (IleRS), and valyl- (ValRS) tRNA synthetases [36, 37]. The class II prolyl-tRNA synthetase (ProRS) possesses a hydrolytic editing domain that can either be inserted or appended to the main body of the enzyme depending on the origins of the ProRS [38]. In bacterial ProRS, the editing domain (INS) is inserted between motifs 2 and 3 and enables ProRS to hydrolyze Ala-tRNAPro [39, 40]. In contrast, ProRS from lower eukaryotes possess an appended editing domain that has moderate homology to the INS [41].
An anticodon binding domain typically provides added specificity for tRNA substrate interactions with the aaRS. With the exception of LeuRS, seryl-tRNA synthetase (SerRS), and alanyl-tRNA synthetase (AlaRS) [1, 42-46], the aaRS interacts with the anticodon loop of tRNA to identify its cognate isoacceptors. Interestingly, co-crystal structures of LeuRS with tRNA indicate that a C-terminal domain extension of LeuRS interacts with the tRNALeu elbow [47, 48]. The anticodon-binding domains are varied in their structure and sequence. For example, those from glutamyl-tRNA synthetase (GluRS) [49] and glutaminyl-tRNA synthetase (GlnRS) [17] are comprised primarily of α-helices and β-strands, respectively. Likewise, sequence alignments of five anticodon binding domains of IleRS ranging from Escherichia coli (E. coli) to humans reveal that they are very divergent depending on their origin [50]. The anticodon binding domains only share a 7% identity residue score that is calculated by dividing the number of identical residues by the total length of the shortest sequence used in the comparison analysis [50].
A small distinctive N-terminal domain in Neurospora crassa (N. crassa) mitochondrial tyrosyl-tRNA synthetase (TyrRS) adds a second essential function to this aaRS that aids in group I intron splicing [51, 52]. The N-terminal region [52, 53] and two short inserts [54] of TyrRS are believed to promote splicing by binding to the group I intron with a newly evolved surface for group I intron binding that is different from where it binds tRNATyr during aminoacylation [55]. In addition, the human TyrRS can also cleave itself to form two distinct cytokines, which are only active after the splitting of TyrRS occurs [56, 57]. The C-terminal domain fragment is conserved with the proinflammatory cytokine known as endothelial-monocyte-activating polypeptide II (EMAP II) [58] and carries out similar immunological activities [57, 59, 60].
An emerging frontier of aaRS research has identified unique non-aminoacylation related domains in eight aaRSs from higher organisms [3]. Some aaRSs idiosyncratically also incorporate known domains including GST [61-67], WHEP [68-75], EMAP II [57, 76-79] and leucine zipper [80-83] domains. Remarkably, the addition of these domains in the tree of life has been correlated with increasing the complexity of organisms and driving the evolution of new more sophisticated species [6].
Amino Acid Fidelity
Accurate discrimination of amino acids by the aaRSs in the first step of protein synthesis is critical to the cell. Charging the wrong amino acid to tRNA would result in statistical mutations in the proteome [84, 85]. Failure to clear these mistakes can result in cell death for microbes and neurodegenerative diseases in mammals [86-89]. Thus, some aaRSs have evolved amino acid editing mechanisms to ensure the highest level of fidelity during protein synthesis [2, 90, 91].
In 1958, Pauling first predicted an error rate of 1 in 20 during protein synthesis when he projected that enzymes would not distinguish fully between structurally similar amino acids that differed by only a single methyl group [92] (Figure 2). Subsequently, Loftfield measured the misincorporation of valine for isoleucine in chicken ovalbumin to be much lower with an error rate of less than 1 in 3000 [93]. It was later found that IleRS could hydrolyze a valyl-adenylate intermediate with the help of tRNA in order to maintain fidelity [94].
Figure 2.
Structurally similar amino acids. A. IleRS discriminates isoleucine from valine. B. AlaRS discriminates alanine from glycine. The colored balls represent as follows: red, oxygen; blue, nitrogen; and gray, carbon. (Hydrogens are not shown.)
Fersht proposed that some aaRSs possess a secondary editing site where error correction promotes higher specificity [95]. In collaboration then, these two separate active sites would form a “double sieve” for amino acid selection to maintain fidelity (Figure 3). The aminoacylation active site would serve as a coarse sieve for amino acid activation in the canonical core, for example, to exclude bulky amino acids, but allow smaller isosteric amino acids to pass through. The second hydrolytic active site operates as a fine sieve to bind to mischarged amino acid for hydrolysis, while blocking correctly charged tRNAs for release to elongation factors. The second sieve was first identified by mutational and deletion analysis of the class I enzyme’s editing domain [96-98]. X-ray crystal structure for the class I IleRS [99] and class II threonyl-tRNA synthetase (ThrRS) [100, 101] showed how the double sieve was separated into two clear protein domains.
Figure 3.
The double sieve of LeuRS. A. The double sieve for LeuRS contains a coarse sieve (red) for aminoacylation that excludes larger amino acids and a “fine sieve” (blue) that blocks cognate amino acid, but allows non-cognate amino acids to be hydrolyzed. B. The crystal structure of T. thermophilus LeuRS has an ancient canonical aminoacylation core (red) and CP1 hydrolytic editing domain (blue) that are linked by β-strands (green). Residues that impact editing within the hydrolytic active site are in orange. Cartoon on the left is adapted from [248] and structure on right (PDB: 2BTE) is adapted from [48].
Most of the editing aaRSs correct amino acid errors in cis via a hydrolytic domain that is inserted into the catalytic core [2]. However, some archaea as well as bacterial aaRSs lack an editing domain that ensures fidelity during protein synthesis. In order to compensate for this, independent paralogs and homologues of editing domains that have hydrolytic activity for amino acid correction have been found to occur naturally. An example is the ProRS editing domain homologue YbaK that edits mischarged Ala-tRNAPro and Cys-tRNAPro in Haemophilus influenzae [40, 102]. Additionally, ProX, a paralog of the ProRS editing domain from Clostridium sticklandii hydrolyzes Ala-tRNAPro [41], while AlaX, an AlaRS homolog from Methanosarcina barkeri and Sulfolobus solfataricus, cleaves mischarged Ser-tRNAAla and Gly-tRNAAla [41, 103].
In either the case of cis or trans editing, the mischarged 3′ end of the tRNA must be transferred from the aminoacylation to the editing active site [104]. Comparison of the crystal structures of LeuRS in various states of binding to tRNA, including an aminoacylation and editing complex of E. coli LeuRS [105] demonstrate that the CP1 domain rotates through the different stages of aminoacylation and its movement is dependent on the flexibility of the β-strands that tether the editing CP1 domain to the main body of the enzyme [2, 48, 106]. Substitution of a specific glycine within the β-strands disrupted tRNA translocation [107]. A “translocation peptide” within the CP1 domain of E. coli LeuRS also facilitates transfer of tRNA from the canonical core to the CP1 domain [104]. Additionally, mutation of a specific residue in IleRS that binds the amino group of mischarged tRNA to the enzyme for editing, prevents translocation of charged tRNA from the aminoacylation active site of the enzyme to its editing pocket [108].
The aaRSs edit by different pathways: pre-transfer and post-transfer of charging to the tRNA. Post-transfer editing hydrolyzes mischarged tRNA to yield free tRNA and non-cognate amino acid (Figure 4) [109]. As discussed above, cleavage of the amino acid from tRNA occurs in the separate hydrolytic editing domain [37, 96] or in some cases via an independent hydrolytic protein [40, 102]. By comparison, pre-transfer editing targets the aminoacyl-adenylate for hydrolysis to release the free amino acid and AMP [94]. Pre-transfer editing can also be dependent on a substrate cyclization reaction such as homocysteine’s conversion to form thiolactone by methionyl-tRNA synthetase (MetRS) [110]. The site where pre-transfer editing occurs is poorly defined and has been hypothesized to be associated with the canonical core [111-114] or the hydrolytic editing domain [25, 108, 115]. A natural LeuRS from Mycoplasma mobile that is completely missing its CP1 domain was shown to have a weak pre-transfer editing activity [116, 117], which was enhanced by a fusion mutant that incorporated a CP1 domain from E. coli. Ironically, CP1 domains originating from ValRS and IleRS, as well as LeuRS enhanced pre-transfer editing activity in M. mobile LeuRS and suggested that this editing domain profoundly influences a pre-transfer editing mechanism that occurs in the LeuRS canonical core [117].
Fig 4.
Pathways for aaRS editing. Post-transfer editing occurs when the incorrectly charged tRNA is hydrolyzed, while pre-transfer editing cleaves activated aminoacyl-adenylate.
Different aaRSs have been suggested to have either pre- or post-transfer editing dominate as the primary fidelity mechanism [95, 118-121]. However, in a backdrop of post-transfer editing, it has been difficult to assess the contribution of pre-transfer editing to achieve amino acid fidelity [120, 122]. In some cases, the aaRS will activate a second editing pathway in the failure of the dominant editing pathway. For example, when the post-transfer editing active site of E. coli LeuRS is removed, the enzyme maintained overall fidelity by activation of a pre-transfer editing pathway [111]. Activation of separate pathways can also be dependent on the amino acid that is misactivated [123, 124]. Specific residues have also been hypothesized to enhance pre-transfer editing [121]. Likewise, when post-transfer editing in the human cytoplasmic LeuRS (hscLeuRS) is mutationally abolished, the enzyme fails to mischarge tRNA, indicating the activation of a second editing pathway to maintain overall accuracy [125].
Cellular mechanisms to ensure amino acid fidelity have been duplicated and even triplicated [2, 120] demonstrating its importance to the cell. Indeed, disruptions to fidelity that lead to even low levels of statistical mutations in the proteome are consequential to the health and viability of cells in both microbial and mammalian systems [84-89]. Yet paradoxically, new examples are emerging, where organisms have naturally evolved to facilitate statistical proteomes. These include Mycoplasma pathogens in which LeuRS, PheRS, and ThrRS contain mutationally disrupted editing domains. M. mobile LeuRS is completely missing its CP1 editing domain [116]. In Candida pathogens, Ser-Leu codon ambiguity persists via tRNA mischarging and also results in statistical proteomes [126, 127]. These examples, suggest that statistical proteomes provide an evolutionary benefit to cells in certain physiological environments.
Aminoacyl-tRNA Synthetase Recognition of tRNA
In general, tRNAs have similar secondary and tertiary structures (Figure 5) [128]. Thus, the aaRSs depend on embedded identity elements within the tRNA substrate to differentiate their cognate substrates for aminoacylation [129-131]. These tRNA identity elements tend to cluster in the anticodon and acceptor stem regions of the tRNA, where there is typically direct interaction with the aaRS [1]. Crystal structures have also identified distal regions, such as the L-shaped tRNA elbow as important recognition sites too [47, 48, 132].
Figure 5.
Secondary and tertiary structure of tRNA. The secondary cloverleaf structure (left) and tertiary structure of tRNA (right) show how the dihydrouridine (D; red) and TΨC (green) loops interact for folding. The colors represented on the tRNA are: orange, 3′ acceptor stem; purple, acceptor stem; green, TΨC stem-loop; red, D stem-loop; light green, variable loop; blue, anticodon stem-loop; and black, anticodon trinucleotide.
Identity elements include single nucleotide variations, such as the discriminator base at position 73 that is located at the end of the tRNA acceptor stem [1, 133]. Based on co-crystal structure information, the tRNA discriminator base is in close proximity to the aaRS active site, providing an opportunity for specific protein-RNA interactions. In addition, unique base pairs such as G3:U70 in the acceptor stem of tRNAAla [129, 134] confer specificity for AlaRS. In other cases, for example tRNASer, the enzyme relies on recognition elements in its D-stem and extra long variable loop [135]. For most aaRSs, distal interactions with the tRNA anticodon nucleotides are required for recognition. The only exceptions are tRNALeu, tRNASer, and tRNAAla [1, 42, 43] as these tRNAs have multiple isoacceptors with too many different anticodon nucleotide combinations to provide a unique anticodon-aaRS match.
Antideterminants prevent productive interactions between the aaRS and noncognate tRNAs. The first antideterminant was identified at position 34 of E. coli tRNAIle, which contains lysidine, a modified cytosine, that blocks misaminoacylation by MetRS [136]. Other examples of antideterminants include modified nucleotide m1G37 in yeast tRNAAsp that hinders aminoacylation by arginyl-tRNA synthetase (ArgRS) [137, 138]. In addition, the G37 nucleotide on tRNASer prevents charging by yeast LeuRS [139]. In contrast, the A73 nucleotide on tRNALeu inhibits aminoacylation by SerRS [140].
A growing base of crystal structures for the family of aaRSs has provided atomic details of the interface between tRNA and the aaRS that facilitate recognition and specificity [1]. However, interpretation using these static structures are limited to specific points of contacts between the protein:RNA pair and fail to provide insight into long-distance effects of distal regions of the tRNA that influence the aaRS active site. A computational approach based on community network analysis has identified communities that are linked by correlated motions and define amino acid and nucleotide nodes that although distal, appear to be critical to tRNA binding and aminoacylation [141].
Macromolecular Complexes of aaRSs Harbor Non-Aminoacylation Activities
In higher eukaryotes, nine aaRSs and three non-synthetase proteins are assembled into a multi-aaRS complex. It was originally hypothesized that assembly of the aaRSs would localize these protein synthesis enzymes for more efficient translation Fox[142, 143]. Recently however, it has been shown that this multi-aaRS complex operates as a “functional depot” for alternate activities of its members [144]. Many of these proteins trigger specific cell signaling activities when freed from the complex.
The macromolecular aaRS complex is comprised of ArgRS, GlnRS, IleRS, LeuRS, LysRS, MetRS, aspartyl- (AspRS), and glutamyl-prolyl-tRNA synthetase (Glu-ProRS) (Figure 7). The Glu-ProRS is an unusual fused aaRS that is linked by three tandem repeating peptides [145, 146]. Three auxiliary non-synthetase proteins called AIMP1, AIMP2, and AIMP3 (aminoacyl-tRNA synthetase interacting multifunctional proteins) formally known as p43, p38, and p18 respectively, are at the core of the multi-aaRS complex [142, 147, 148] (Figure 7).
Figure 7.
Organization of the multi-aaRS complex. The aaRS complex consists of nine aaRSs (gray) and three auxiliary proteins (green). The varied sizes of the proteins are schematically indicated by different sized balls. Dashed lines indicate sub-complexes. Solid lines represent protein-protein interactions with the exception of Glu-ProRS, which is covalently fused by a repeating peptide motif.
A 30 Å cryo-electron microscopy (EM) structure of the multi-aaRS complex is shaped as an asymmetric triangle with dimensions of approximately 19 × 16 × 10 nm. A central cleft that is 4 nm deep extends about two-thirds of the length of the complex [149, 150]. Biochemical investigations indicate that the large complex can be partitioned into three sub-complexes. Based on tandem affinity purification from human cells, AIMP2 interacts with AspRS and LysRS and provides a foundation for complex assembly [66, 80] by connecting the other two complexes [147, 151]. The AIMP1 protein interacts with GlnRS and the N-terminus of ArgRS to hold them together with the rest of the complex [147, 152, 153]. The AIMP3 anchors MetRS and also interacts with the other multi-aaRS components such as Glu-ProRS, IleRS, and LeuRS [147].
Certain events prompt the release of individual aaRSs or one of the AIMP proteins to elicit a specific activity or cell signaling cascade [154] (Table 2). In at least one case, phosphorylation was critical to the release of AIMP2 from the multi-synthetase complex in response to DNA damage [155]. Phosphorylation of AIMP1 is also important to its regulation of an immune reaction [156, 157].
Table 2.
Representative activities associated with aaRSs and AIMPs when free of the human macromolecular complex
| Proteins | Activities |
|---|---|
| LysRS-II | Inflammatory cytokine (Extracellular) Viral assembly (Plasma membrane) Transcriptional control (Nuclear) |
| Glu-ProRS | Translational silencing |
| GlnRS | Anti-apoptosis |
| MetRS | rRNA transcription |
| IleRS | Autoimmune response |
| AIMP1 | Cytokine activation of macrophages |
| AIMP2 | Degradation of FBP, a transcriptional activator of c- myc |
| AIMP3 | p53 induction and DNA repair by activating ATM/ATR in the nucleus |
In human cells, LysRS generates diadenosine tetraphosphate (Ap4A), a critical signalling molecule, in order to activate mast cells. The Ap4A binds to the tumor suppressor gene Hint to decrease interactions between Hint and the transcription factor MITF [158]. This causes MITF to dissociate from LysRS as well as Hint, thus releasing it to transactivate its responsive genes [158]. The LysRS can also act as a pro-inflammatory cytokine [159] if its release is induced by tumor necrosis factor α (TNF-α). This secreted LysRS can bind macrophages in order to enhance TNF-α production and migration in a positive feedback loop [159, 160].
Glu-ProRS can be switched from its classical role in translation by phosphorylation of one of the linking WHEP domains to become a gene-specific modulator of translation [161]. It forms a complex called GAIT (gamma-interferon-activated inhibitor of translation) with ribosomal subunit L13a and glyceraldehyde-3-phosphate (GAPDH) in response to interferon-γ to trigger translational silencing in ceruloplasmin. The Glu-ProRS complex binds to the 3′-untranslated region of ceruloplasmin to block mRNA translation [162]. Specific translational silencing of ceruloplasmin has been hypothesized to serve as a feedback mechanism to prevent over-accumulation of the protein, which might result in oxidative damage [162].
Lower eukaryotes such as S. cerevisiae also contain multi-synthetase complexes although comprising fewer components than those found in mammals [163]. In the case of S. cerevisiae, Arc1p binds to tRNA and associates with MetRS as well as GluRS in order to function as a co-factor for these two aaRSs [163, 164]. Similarly, the archaea Methanothermobacter thermautotrophicus harbors a small multi-synthetase complex comprising LeuRS, LysRS and ProRS that enhances tRNAPro aminoacylation [165, 166]. In addition, kinetic experiments indicate that the catalytic turnovers of LysRS and ProRS increased when in complex as compared to the free enzyme [167].
The three auxiliary proteins, AIMP1, AIMP2, and AIMP3 associated with the macromolecular aaRS complex in the cell possess signaling activities when freed from the complex. For example, AIMP1 is secreted as a cytokine and stimulates fibroblast proliferation [168]. It also has pro-apoptotic function [169], and is involved in angiogenesis [170]. AIMP2 displays pro-apoptotic function [171] and is the target of parkin for ubiquitination, leading to neurodegeneration [172]. Finally, AIMP3 acts as a tumor suppressor by activating ataxia-telangiectasia, mutated/ATM and Rad3-related (ATM/ATR) which is an upstream kinase of p53 [173]. This is essential for subsequent p53 induction and DNA repair [64, 173].
Secondary and Alternate Functions of aaRSs
The aaRSs are considered to be one of the most ancient families of proteins. Throughout their long evolutionary history, these enzymes have adapted to carry out dual roles in the cell that include non-protein synthesis activities (Figure 6) [174]. In some cases, aaRS-like proteins or paralogs of aaRSs that originate from complete or partial gene duplications of the aaRS gene have also adapted for diverse cellular functions [5] (Figure 6). A number of these alternate functions have been implicated in disease states [175]. They have also been proposed to drive organism evolution by conferring increased complexity [6].
Figure 6.
Non-canonical activities and paralogs of aaRSs. A. The aaRSs are adapted for dual roles that coexist with their aminoacylation activity. B. Paralogs of aaRSs and their domains can provide important non-aminoacylation functions within the cell [183].
The secondary activities of aaRSs capitalize on their existing binding sites for RNA, amino acid and ATP [3]. They also take advantage of inserted domains or peptides to adapt for alternate activities [3]. As Figure 6 shows, these alternate roles are broad in scope and include mitochondrial RNA splicing [51, 176-179], immunology [56, 57, 154], translational [180], and transcriptional regulation [181, 182]. In many cases, the alternate function is masked and the aaRS requires modification and/or proteolysis to unleash the unconventional activity [144]. Some disease states including neurodegenerative and autoimmune disorders have implicated or identified an aaRS alternate functions when it has been compromised or failed [183].
The glycyl-tRNA synthetase (GlyRS) [184] as well as TyrRS [185] have been associated with Charcot-Marie-Tooth (CMT) disease which is a heritable disorder of the peripheral nervous system [186]. Interestingly, CMT-causing mutations lie at the dimer interface of GlyRS, and presumably affect interactions between the two subunits [187]. In addition, mutations of a highly conserved glycine (G41R) involved in tyrosyl-adenylate formation as well as the Glu196 residue in human TyrRS that make contact with tRNATyr result in this disease phenotypes too [185]. While the mechanism of GlyRS and TyrRS involvement remains unclear, investigations with differentiating motor neurons indicate that wild type TyrRS localizes to the axonal termini. In contrast, in the case of neurons expressing mutant TyrRS proteins, their normal function and axonal distribution is disrupted leading to axonal loss, degeneration and peripheral neuropathy.
Some auto-immune disorders that are associated with aaRSs include chronic myopathies and interstitial lung diseases. For example, auto-antibodies against HisRS are associated with chronic muscle inflammatory disease or polymyositis a form of idiopathic inflammatory myopathy [188, 189]. Antibodies to AsnRS are linked to auto-immune interstitial lung disease [190, 191]. IleRS and GlyRS were also shown to be antigenic in screens for auto-antibodies generated in myositis patients [192-194].
Many of the aaRSs or a fragment of the aaRS are involved in cell signaling activities. This includes an immunology activity during apoptosis for eukaryotic TyrRS [56, 57], when it is cleaved into its N- and C- terminal halves to yield two distinct cytokines. The TyrRS C-terminal fragment up-regulates TNF-α, while its N-terminal fragment mimics the functions of interleukin-8 (IL-8) that is released by macrophages in an antigen response [56, 57]. In the case of parasite protozoan Entamoeba histolytica, LysRS is up-regulated by TNF-α and then is cleaved by proteases to release a cytokine-like domain that functions in chemotaxis [195]. A TrpRS signal transducer activity mediates interactions between DNA-dependent protein kinases and poly ADP ribose polymerase leading to downstream signaling events for p53 activation [196]. GlyRS can be secreted from macrophages and exhibits cytokine-like properties that down-regulate ERK MAP kinases to inhibit the Ras activated tumor cells. In particular, GlyRS binding to CDH6 (cadherin) releases suppressed phosphatase 2A (PP2A) that dephosphorylates activated ERK [197]. As another example, the HisRS amino terminal domain functions as a chemokine through CCR5 receptor-mediated interactions [198].
As introduced above, the eukaryotic multi-synthetase complex has been termed a depot that sequesters signaling functions for these housekeeping proteins [144]. For example, when LysRS, is released from the complex upon MAPK-dependent phosphorylation, it enters the nucleus to activate transcription of immune response genes [154]. Then, in response to immunogenic cell death inducers, LysRS translocates to the cell surface along with calrectulin that acts as an engulfment signal for antigen-presenting cells [199]. In a second example, Glu-ProRS is phosphorylated by Cdk5, which is responsible for translational control of macrophage inflammatory gene expression [200]. Additionally Glu-ProRS also is involved in the translational silencing of mRNAs encoding vascular endothelial growth factor A (VEGF-A). Under INF-γ stimulus, the phosphorylated Glu-ProRS is recruited to the GAIT complex for the translational silencing event. However, VEGF-A is required for vessel maintenance and in order to maintain its basal expression, a C-terminally truncated form of Glu-ProRS protects the mRNA from the GAIT complex assembly [201].
AsnRS was found to be one of the important components of survival signaling cascade induced by fibroblast growth factor 2(FGF2) in osteoblasts [202]. In addition, filarial parasite Brugia malayi AsnRS interacts with and activates human IL-8 cytokine receptors CXCR1 and CXCR2. It also stimulates MAP kinases like ERK1 and ERK2 in cell migration and signal transduction [203].
In lower eukaryotes, two different mitochondrial aaRSs adapted to support RNA splicing, but via distinct mechanisms. The N. crassa TyrRS has acquired a small idiosyncratic N-terminal α-helical domain [52] as well as two short inserts [54] for splicing activity. Its X-ray co-crystal structure shows that TyrRS binds to the group I intron with a newly evolved surface for group I intron binding that is different from where it binds tRNATyr during aminoacylation [55]. In contrast, Saccharomyces cerevisiae (S. cerevisiae) mitochondrial LeuRS appears to require no special adaptations to confer RNA splicing of group I introns [204]. The CP1 editing domain is important to its activity and can function in splicing independent of the full-length enzyme [205]. In both examples of aaRS-dependent group I intron splicing however, it is proposed that they bind introns and promote folding of their cognate introns into an active conformation [205-207].
In E. coli, regulation of gene expression at transcription and translational levels is aided in part by AlaRS and ThrRS. Biochemical studies showed that AlaRS binds to a palindromic sequence that flanks its promoter site to repress transcription of its own gene [182]. Also, ThrRS regulates its production at the translational level by preventing ribosome binding for translation [180]. The enzyme binds to two stem-loop structures upstream of the ribosome binding site, which mimic the anticodon arm of tRNAThr [180].
Human LysRS is packaged into human immunodeficiency virus type 1 (HIV-1) together with its cognate tRNALys [208, 209]. Packaging depends on a specific interaction of LysRS with the viral Gag protein [210, 211] and Gag-Pol has also been implicated in LysRS recruitment and packaging [212-215]. Interestingly, the binding of tRNALys3 to LysRS rather than aminoacylation of tRNALys3 is critical for tRNALys3 packaging into HIV [208, 212]. Once tRNALys3 is inside the virion, it anneals to the genomic RNA of the virus and primes reverse transcription [216].
Several aaRSs charge non-canonical tRNA substrates that have unusual functions. For example, E. coli AlaRS aminoacylates transfer-messenger RNA (tmRNA). The charged tmRNA binds to stalled ribosomes and provides its own template to translate a specific C-terminal fusion peptide that targets the truncated peptide for degradation [217]. Additionally, aminoacylated tRNAs can be used for peptide cell wall biosynthesis. For example, a species of tRNAGly [218] is involved in peptidoglycan synthesis, while Lys-tRNALys can be used to remodel lipids [219].
Some plant RNA viral genomes possess tRNA like structures (TLS) [220] at their 3′ ends that can be specifically aminoacylated by valine [221, 222], histidine [223-225] or tyrosine [226, 227]. The TLSs lack the modified bases, conserved residues, D loop and T loop sequences characteristic of tRNAs [228]. However, the TLSs contain a pseudo-knotted aminoacyl acceptor stem which places the 5′ end of the TLS away from the acceptor end [229, 230]. Interaction of the TLS with the tRNA synthetase predominantly occurs via the anticodon and the acceptor arms despite the presence of the pseudoknot [227, 231].
Paralogs of aaRSs or aaRS-like proteins have been identified that take part in diverse cellular activities (Figure 6) [5]. For instance, E. coli YadB protein resembles GluRS and attaches glutamate to a queuosine base (position 34) at the first anticodon position of tRNAAsp [232]. This modification expands and alters tRNA decoding of codons [232, 233]. HisZ, a paralog of HisRS, is one component of the enzyme catalyzing the first step of histidine biosynthesis [234]. As indicated above, aaRSs that lack an editing domain such as ProRS [40] and AlaRS [103], have paralogs and homologs of editing domains in order to maintain overall fidelity [41].
aaRSs as Drug Targets
As essential enzymes, the aaRSs are ideal targets for antibiotics, with the caveat that these drugs must be species-specific to avoid inhibiting the human counterpart. The most successful example is mupirocin (pseudomonic acid A), a topical antibiotic for Staphylococcus aureus [235] that inhibits IleRS [236, 237]. It functions by interfering with isoleucine as well as ATP binding to IleRS [238, 239]. Mutational analysis has identified two residues in the IleRS active site that play a significant role in conferring mupirocin resistance [25].
More recently, a small boron-containing molecule called AN2690 (5-fluoro-1,3-dihydroxy-2,1-benzoxaborole) was discovered that targets the editing active site of LeuRS [240] (Figure 8). The mechanism of action for AN2690 relies on the boron to trap tRNALeu within the editing active site of LeuRS. AN2690 binds specifically to the editing site and the boron attacks the ribose cis diols at the 3′ end of the tRNA to form an oxaborole adduct. This potent therapeutic kills a range of organisms including yeasts, molds and dermatophytes [241] and has just emerged from clinical trials for commercialization to treat onychomycosis.
Figure 8.
Structure of boron-containing AN2690
Other potential therapeutics involving aaRSs include a natural fragment of the TyrRS enzyme (mini TyrRS) that contains the aminoacylation catalytic domain. It is involved in angiogenesis by functioning as a proangiogenic cytokine [242]. Mini-TyrRS binds to endothelial cells and activates an angiogenic signaling pathway that can potentially be harnessed to combat cardiovascular disease [242]. In addition, an alternative splice fragment of TrpRS [243] as well as a natural proteolytic fragment (T2-TrpRS) [244] have strong anti-angiogenic activity and function by inhibiting vascular endothelial growth factor-induced angiogenesis [244]. The aminoacylation active site of T2-TrpRS interacts with two tryptophan residues of VE-cadherin which is a main constituent of adherens (protein complexes at cellular junctions) junctions of endothelial cells [245]. During new blood vessel formation, T2-TrpRS exhibits anti-angiogeneic activity by binding to the exposed Trp residues on VE-cadherin unlike in the case of developed blood vessels where the Trp2 and Trp4 of VE-cadherins are interlocked, buried and unavailable for T2-TrpRS binding. Anti-angiogenic therapy studies suggest that TrpRS is a promising natural therapeutic to prevent blindness caused by abnormal angiogenesis in the eye [246].
Summary and Perspective
The aaRSs have long been known for their prominent role in protein synthesis and have served as an extraordinary model to investigate the rules that govern RNA-protein interactions, enzyme catalysis, protein structure-function relationships, and the principles of evolution. Recent advances suggest that they will also offer profound insight into the functional expansion of the proteome and systems biology as the depths of the aaRSs’ alternate activities are further unraveled and characterized. It is clear that nature has repeatedly called upon the aaRSs throughout the evolution of the cell to adapt as the physiology and metabolic pathways of its environment became more sophisticated. Many of the resultant novel complexities and idiosyncracies that are associated with each of the members of the aaRS family have potential for development into therapeutics to combat human diseases.
Acknowledgements
This work was supported by grants from the National Institutes of Health (GM63789) and the National Science Foundation (MCB0843611). We thank Professors S. Kim, K. Musier-Forsyth, and M. Ibba for providing feedback on parts of our manuscript as well as Hanchao Zhao for technical assistance.
References
- 1.Giegé R, Sissler M, Florentz C. Universal rules and idiosyncratic features in tRNA identity. Nucl Acids Res. 1998;26(22):5017–5035. doi: 10.1093/nar/26.22.5017. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 2.Mascarenhas AP, An S, Rosen AE, Martinis SA, Musier-Forsyth K. Fidelity mechanisms of the aminoacyl-tRNA synthetases. In: RajBhandary UL, Köhrer C, editors. Protein Engineering. Springer-Verlag New York Inc.; New York: 2008. [Google Scholar]
- 3.Guo M, Schimmel P, Yang XL. Functional expansion of human tRNA synthetases achieved by structural inventions. FEBS Lett. 2010;584(2):434–442. doi: 10.1016/j.febslet.2009.11.064. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 4.Nawaz M, Martinis SA. Encyclopedia of Chemical Biology. John Wiley & Sons; New York: 2009. Chemistry of aminoacyl-tRNA synthetases. [Google Scholar]
- 5.Schimmel P, Ribas De Pouplana L. Footprints of aminoacyl-tRNA synthetases are everywhere. Trends Biochem Sci. 2000;25(5):207–209. doi: 10.1016/s0968-0004(00)01553-x. [DOI] [PubMed] [Google Scholar]
- 6.Guo M, Yang XL, Schimmel P. New functions of aminoacyl-tRNA synthetases beyond translation. Nature Rev. 2010;11(9):668–674. doi: 10.1038/nrm2956. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 7.Cusack S, Berthet-Colominas C, Härtlein M, Nassar N, Leberman R. A second class of synthetase structure revealed by X-ray analysis of Escherichia coli seryl-tRNA synthetase at 2.5 Å. Nature. 1990;347(6290):249–255. doi: 10.1038/347249a0. [DOI] [PubMed] [Google Scholar]
- 8.Eriani G, Delarue M, Poch O, Gangloff J, Moras D. Partition of tRNA synthetases into two classes based on mutually exclusive sets of sequence motifs. Nature. 1990;347(6289):203–206. doi: 10.1038/347203a0. [DOI] [PubMed] [Google Scholar]
- 9.Ribas de Pouplana L, Schimmel P. Two classes of tRNA synthetases suggested by sterically compatible dockings on tRNA acceptor stem. Cell. 2001;104(2):191–193. doi: 10.1016/s0092-8674(01)00204-5. [DOI] [PubMed] [Google Scholar]
- 10.Burbaum JJ, Schimmel P. Assembly of a class I tRNA synthetase from products of an artificially split gene. Biochemistry. 1991;30(2):319–324. doi: 10.1021/bi00216a002. [DOI] [PubMed] [Google Scholar]
- 11.Cusack S. Eleven down and nine to go. Nature Struct Biol. 1995;2(10):824–831. doi: 10.1038/nsb1095-824. [DOI] [PubMed] [Google Scholar]
- 12.Landes C, Perona JJ, Brunie S, Rould MA, Zelwer C, Steitz TA, Risler JL. A structure-based multiple sequence alignment of all class I aminoacyl-tRNA synthetases. Biochimie. 1995;77(3):194–203. doi: 10.1016/0300-9084(96)88125-9. [DOI] [PubMed] [Google Scholar]
- 13.Moras D. Structural and functional relationships between aminoacyl-tRNA synthetases. Trends Biochem Sci. 1992;17(4):159–164. doi: 10.1016/0968-0004(92)90326-5. [DOI] [PubMed] [Google Scholar]
- 14.Ibba M, Morgan S, Curnow AW, Pridmore DR, Vothknecht UC, Gardner W, Lin W, Woese CR, Söll D. A euryarchaeal lysyl-tRNA synthetase: resemblance to class I synthetases. Science. 1997;278(5340):1119–1122. doi: 10.1126/science.278.5340.1119. [DOI] [PubMed] [Google Scholar]
- 15.Irwin MJ, Nyborg J, Reid BR, Blow DM. The crystal structure of tyrosyl-transfer RNA synthetase at 2.7 Å resolution. J Mol Biol. 1976;105(4):577–586. doi: 10.1016/0022-2836(76)90236-9. [DOI] [PubMed] [Google Scholar]
- 16.Brick P, Bhat TN, Blow DM. Structure of tyrosyl-tRNA synthetase refined at 2.3 Å resolution. Interaction of the enzyme with the tyrosyl adenylate intermediate. J Mol Biol. 1989;208(1):83–98. doi: 10.1016/0022-2836(89)90090-9. [DOI] [PubMed] [Google Scholar]
- 17.Rould MA, Perona JJ, Soll D, Steitz TA. Structure of E. coli glutaminyl-tRNA synthetase complexed with tRNA(Gln) and ATP at 2.8 Å resolution. Science. 1989;246(4934):1135–1142. doi: 10.1126/science.2479982. [DOI] [PubMed] [Google Scholar]
- 18.Schmitt E, Meinnel T, Blanquet S, Mechulam Y. Methionyl-tRNA synthetase needs an intact and mobile 332KMSKS336 motif in catalysis of methionyl adenylate formation. J Mol Biol. 1994;242(4):566–576. doi: 10.1006/jmbi.1994.1601. [DOI] [PubMed] [Google Scholar]
- 19.Webster T, Tsai H, Kula M, Mackie GA, Schimmel P. Specific sequence homology and three-dimensional structure of an aminoacyl transfer RNA synthetase. Science. 1984;226(4680):1315–1317. doi: 10.1126/science.6390679. [DOI] [PubMed] [Google Scholar]
- 20.Perona JJ, Rould MA, Steitz TA. Structural basis for transfer RNA aminoacylation by Escherichia coli glutaminyl-tRNA synthetase. Biochemistry. 1993;32(34):8758–8771. doi: 10.1021/bi00085a006. [DOI] [PubMed] [Google Scholar]
- 21.Arnez JG, Moras D. Structural and functional considerations of the aminoacylation reaction. Trends Biochem Sci. 1997;22(6):211–216. doi: 10.1016/s0968-0004(97)01052-9. [DOI] [PubMed] [Google Scholar]
- 22.Fersht AR, Knill-Jones JW, Bedouelle H, Winter G. Reconstruction by site-directed mutagenesis of the transition state for the activation of tyrosine by the tyrosyl-tRNA synthetase: a mobile loop envelopes the transition state in an induced-fit mechanism. Biochemistry. 1988;27(5):1581–1587. doi: 10.1021/bi00405a028. [DOI] [PubMed] [Google Scholar]
- 23.First EA, Fersht AR. Analysis of the role of the KMSKS loop in the catalytic mechanism of the tyrosyl-tRNA synthetase using multimutant cycles. Biochemistry. 1995;34(15):5030–5043. doi: 10.1021/bi00015a014. [DOI] [PubMed] [Google Scholar]
- 24.Ilyin VA, Temple B, Hu M, Li G, Yin Y, Vachette P, Carter CW., Jr. 2.9 Å crystal structure of ligand-free tryptophanyl-tRNA synthetase: domain movements fragment the adenine nucleotide binding site. Protein Sci. 2000;9(2):218–231. doi: 10.1110/ps.9.2.218. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 25.Nakama T, Nureki O, Yokoyama S. Structural basis for the recognition of isoleucyl-adenylate and an antibiotic, mupirocin, by isoleucyl-tRNA synthetase. J Biol Chem. 2001;276(50):47387–47393. doi: 10.1074/jbc.M109089200. [DOI] [PubMed] [Google Scholar]
- 26.Rossmann MG, Moras D, Olsen KW. Chemical and biological evolution of nucleotide-binding protein. Nature. 1974;250(463):194–199. doi: 10.1038/250194a0. [DOI] [PubMed] [Google Scholar]
- 27.Arnez JG, Harris DC, Mitschler A, Rees B, Francklyn CS, Moras D. Crystal structure of histidyl-tRNA synthetase from Escherichia coli complexed with histidyl-adenylate. EMBO J. 1995;14(17):4143–4155. doi: 10.1002/j.1460-2075.1995.tb00088.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 28.Ruff M, Krishnaswamy S, Boeglin M, Poterszman A, Mitschler A, Podjarny A, Rees B, Thierry JC, Moras D. Class II aminoacyl transfer RNA synthetases: crystal structure of yeast aspartyl-tRNA synthetase complexed with tRNA(Asp) Science. 1991;252(5013):1682–1689. doi: 10.1126/science.2047877. [DOI] [PubMed] [Google Scholar]
- 29.O’Donoghue P, Sethi A, Woese CR, Luthey-Schulten ZA. The evolutionary history of Cys-tRNACys formation. Proc Natl Acad Sci U S A. 2005;102(52):19003–19008. doi: 10.1073/pnas.0509617102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 30.Polycarpo C, Ambrogelly A, Berube A, Winbush SM, McCloskey JA, Crain PF, Wood JL, Söll D. An aminoacyl-tRNA synthetase that specifically activates pyrrolysine. Proc Natl Acad Sci U S A. 2004;101(34):12450–12454. doi: 10.1073/pnas.0405362101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 31.Polycarpo C, Ambrogelly A, Ruan B, Tumbula-Hansen D, Ataide SF, Ishitani R, Yokoyama S, Nureki O, Ibba M, Söll D. Activation of the pyrrolysine suppressor tRNA requires formation of a ternary complex with class I and class II lysyl-tRNA synthetases. Mol Cell. 2003;12(2):287–294. doi: 10.1016/s1097-2765(03)00280-6. [DOI] [PubMed] [Google Scholar]
- 32.Brick P, Blow DM. Crystal structure of a deletion mutant of a tyrosyl-tRNA synthetase complexed with tyrosine. J Mol Biol. 1987;194(2):287–297. doi: 10.1016/0022-2836(87)90376-7. [DOI] [PubMed] [Google Scholar]
- 33.Belrhali H, Yaremchuk A, Tukalo M, Larsen K, Berthet-Colominas C, Leberman R, Beijer B, Sproat B, Als-Nielsen J, Grubel G, et al. Crystal structures at 2.5 Å resolution of seryl-tRNA synthetase complexed with two analogs of seryl adenylate. Science. 1994;263(5152):1432–1436. doi: 10.1126/science.8128224. [DOI] [PubMed] [Google Scholar]
- 34.Sprinzl M, Cramer F. Site of aminoacylation of tRNAs from Escherichia coli with respect to the 2′- or 3′-hydroxyl group of the terminal adenosine. Proc Natl Acad Sci U S A. 1975;72(8):3049–3053. doi: 10.1073/pnas.72.8.3049. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 35.Starzyk RM, Webster TA, Schimmel P. Evidence for dispensable sequences inserted into a nucleotide fold. Science. 1987;237(4822):1614–1618. doi: 10.1126/science.3306924. [DOI] [PubMed] [Google Scholar]
- 36.Betha AK, Williams AM, Martinis SA. Isolated CP1 domain of Escherichia coli leucyl-tRNA synthetase is dependent on flanking hinge motifs for amino acid editing activity. Biochemistry. 2007;46(21):6258–6267. doi: 10.1021/bi061965j. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 37.Lin L, Hale SP, Schimmel P. Aminoacylation error correction. Nature. 1996;384(6604):33–34. doi: 10.1038/384033b0. [DOI] [PubMed] [Google Scholar]
- 38.Beuning PJ, Musier-Forsyth K. Hydrolytic editing by a class II aminoacyl-tRNA synthetase. Proc Natl Acad Sci U S A. 2000;97(16):8916–8920. doi: 10.1073/pnas.97.16.8916. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 39.Wong FC, Beuning PJ, Nagan M, Shiba K, Musier-Forsyth K. Functional role of the prokaryotic proline-tRNA synthetase insertion domain in amino acid editing. Biochemistry. 2002;41(22):7108–7115. doi: 10.1021/bi012178j. [DOI] [PubMed] [Google Scholar]
- 40.Wong FC, Beuning PJ, Silvers C, Musier-Forsyth K. An isolated class II aminoacyl-tRNA synthetase insertion domain is functional in amino acid editing. J Biol Chem. 2003;278(52):52857–52864. doi: 10.1074/jbc.M309627200. [DOI] [PubMed] [Google Scholar]
- 41.Ahel I, Korencic D, Ibba M, Söll D. Trans-editing of mischarged tRNAs. Proc Natl Acad Sci U S A. 2003;100(26):15422–15427. doi: 10.1073/pnas.2136934100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 42.Asahara H, Himeno H, Tamura K, Hasegawa T, Watanabe K, Shimizu M. Recognition nucleotides of Escherichia coli tRNALeu and its elements facilitating discrimination from tRNASer and tRNATyr. J Mol Biol. 1993;231(2):219–229. doi: 10.1006/jmbi.1993.1277. [DOI] [PubMed] [Google Scholar]
- 43.Asahara H, Himeno H, Tamura K, Nameki N, Hasegawa T, Shimizu M. Discrimination among E. coli tRNAs with a long variable arm. Nucl Acids Symp Ser. 1993;(29):207–208. [PubMed] [Google Scholar]
- 44.Dock-Bregeon AC, Garcia A, Giegé R, Moras D. The contacts of yeast tRNASer with seryl-tRNA synthetase studied by footprinting experiments. Eur J Bioch. 1990;188(2):283–290. doi: 10.1111/j.1432-1033.1990.tb15401.x. [DOI] [PubMed] [Google Scholar]
- 45.Normanly J, Ogden RC, Horvath SJ, Abelson J. Changing the identity of a transfer RNA. Nature. 1986;321(6067):213–219. doi: 10.1038/321213a0. [DOI] [PubMed] [Google Scholar]
- 46.Park SJ, Schimmel P. Evidence for interaction of an aminoacyl transfer RNA synthetase with a region important for the identity of its cognate transfer RNA. J Biol Chem. 1988;263(32):16527–16530. [PubMed] [Google Scholar]
- 47.Hsu JL, Martinis SA. A Flexible peptide tether controls accessibility of a unique C-terminal RNA-binding domain in leucyl-tRNA synthetases. J Mol Biol. 2008;376(2):482–491. doi: 10.1016/j.jmb.2007.11.065. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 48.Tukalo M, Yaremchuk A, Fukunaga R, Yokoyama S, Cusack S. The crystal structure of leucyl-tRNA synthetase complexed with tRNALeu in the post-transfer-editing conformation. Nature Struct Mol Biol. 2005;12(10):923–930. doi: 10.1038/nsmb986. [DOI] [PubMed] [Google Scholar]
- 49.Nureki O, Vassylyev DG, Katayanagi K, Shimizu T, Sekine S, Kigawa T, Miyazawa T, Yokoyama S, Morikawa K. Architectures of class-defining and specific domains of glutamyl-tRNA synthetase. Science. 1995;267(5206):1958–1965. doi: 10.1126/science.7701318. [DOI] [PubMed] [Google Scholar]
- 50.Shiba K, Suzuki N, Shigesada K, Namba Y, Schimmel P, Noda T. Human cytoplasmic isoleucyl-tRNA synthetase: selective divergence of the anticodon-binding domain and acquisition of a new structural unit. Proc Natl Acad Sci U S A. 1994;91(16):7435–7439. doi: 10.1073/pnas.91.16.7435. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 51.Akins RA, Lambowitz AM. A protein required for splicing group I introns in Neurospora mitochondria is mitochondrial tyrosyl-tRNA synthetase or a derivative thereof. Cell. 1987;50(3):331–345. doi: 10.1016/0092-8674(87)90488-0. [DOI] [PubMed] [Google Scholar]
- 52.Cherniack AD, Garriga G, Kittle JD, Jr., Akins RA, Lambowitz AM. Function of Neurospora mitochondrial tyrosyl-tRNA synthetase in RNA splicing requires an idiosyncratic domain not found in other synthetases. Cell. 1990;62(4):745–755. doi: 10.1016/0092-8674(90)90119-y. [DOI] [PubMed] [Google Scholar]
- 53.Mohr G, Rennard R, Cherniack AD, Stryker J, Lambowitz AM. Function of the Neurospora crassa mitochondrial tyrosyl-tRNA synthetase in RNA splicing. Role of the idiosyncratic N-terminal extension and different modes of interaction with different group I introns. J Mol Biol. 2001;307(1):75–92. doi: 10.1006/jmbi.2000.4460. [DOI] [PubMed] [Google Scholar]
- 54.Paukstelis PJ, Coon R, Madabusi L, Nowakowski J, Monzingo A, Robertus J, Lambowitz AM. A tyrosyl-tRNA synthetase adapted to function in group I intron splicing by acquiring a new RNA binding surface. Mol Cell. 2005;17(3):417–428. doi: 10.1016/j.molcel.2004.12.026. [DOI] [PubMed] [Google Scholar]
- 55.Paukstelis PJ, Chen JH, Chase E, Lambowitz AM, Golden BL. Structure of a tyrosyl-tRNA synthetase splicing factor bound to a group I intron RNA. Nature. 2008;451(7174):94–97. doi: 10.1038/nature06413. [DOI] [PubMed] [Google Scholar]
- 56.Wakasugi K, Schimmel P. Highly differentiated motifs responsible for two cytokine activities of a split human tRNA synthetase. J Biol Chem. 1999;274(33):23155–23159. doi: 10.1074/jbc.274.33.23155. [DOI] [PubMed] [Google Scholar]
- 57.Wakasugi K, Schimmel P. Two distinct cytokines released from a human aminoacyl-tRNA synthetase. Science. 1999;284(5411):147–151. doi: 10.1126/science.284.5411.147. [DOI] [PubMed] [Google Scholar]
- 58.Kleeman TA, Wei D, Simpson KL, First EA. Human tyrosyl-tRNA synthetase shares amino acid sequence homology with a putative cytokine. J Biol Chem. 1997;272(22):14420–14425. doi: 10.1074/jbc.272.22.14420. [DOI] [PubMed] [Google Scholar]
- 59.Kao J, Houck K, Fan Y, Haehnel I, Libutti SK, Kayton ML, Grikscheit T, Chabot J, Nowygrod R, Greenberg S, Kuang W-J, Leung DW, Hayward JR, Kisiell W, Heath M, Brett J, Stern DM. Characterization of a novel tumor-derived cytokine. Endothelial-monocyte activating polypeptide II. J Biol Chem. 1994;269(40):25106–25119. [PubMed] [Google Scholar]
- 60.Kao J, Ryan J, Brett G, Chen J, Shen H, Fan YG, Godman G, Familletti PC, Wang F, Pan Y-C, Stern D, Clauss M. Endothelial monocyte-activating polypeptide II. A novel tumor-derived polypeptide that activates host-response mechanisms. J Biol Chem. 1992;267(28):20239–20247. [PubMed] [Google Scholar]
- 61.Bec G, Kerjan P, Waller JP. Reconstitution in vitro of the valyl-tRNA synthetase-elongation factor (EF) 1 beta gamma delta complex. Essential roles of the NH2-terminal extension of valyl-tRNA synthetase and of the EF-1 delta subunit in complex formation. J Biol Chem. 1994;269(3):2086–2092. [PubMed] [Google Scholar]
- 62.Jiang S, Wolfe CL, Warrington JA, Norcum MT. Three-dimensional reconstruction of the valyl-tRNA synthetase/elongation factor-1H complex and localization of the delta subunit. FEBS Lett. 2005;579(27):6049–6054. doi: 10.1016/j.febslet.2005.09.062. [DOI] [PubMed] [Google Scholar]
- 63.Kim JE, Kim KH, Lee SW, Seol W, Shiba K, Kim S. An elongation factor-associating domain is inserted into human cysteinyl-tRNA synthetase by alternative splicing. Nucl Acids Res. 2000;28(15):2866–2872. doi: 10.1093/nar/28.15.2866. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 64.Kim KJ, Park MC, Choi SJ, Oh YS, Choi EC, Cho HJ, Kim MH, Kim SH, Kim DW, Kim S, et al. Determination of three-dimensional structure and residues of the novel tumor suppressor AIMP3/p18 required for the interaction with ATM. J Biol Chem. 2008;283(20):14032–14040. doi: 10.1074/jbc.M800859200. [DOI] [PubMed] [Google Scholar]
- 65.Negrutskii BS, Shalak VF, Kerjan P, El’skaya AV, Mirande M. Functional interaction of mammalian valyl-tRNA synthetase with elongation factor EF-1α in the complex with EF-1H. J Biol Chem. 1999;274(8):4545–4550. doi: 10.1074/jbc.274.8.4545. [DOI] [PubMed] [Google Scholar]
- 66.Quevillon S, Robinson JC, Berthonneau E, Siatecka M, Mirande M. Macromolecular assemblage of aminoacyl-tRNA synthetases: identification of protein-protein interactions and characterization of a core protein. J Mol Biol. 1999;285(1):183–195. doi: 10.1006/jmbi.1998.2316. [DOI] [PubMed] [Google Scholar]
- 67.Simader H, Hothorn M, Kohler C, Basquin J, Simos G, Suck D. Structural basis of yeast aminoacyl-tRNA synthetase complex formation revealed by crystal structures of two binary sub-complexes. Nucl Acids Res. 2006;34(14):3968–3979. doi: 10.1093/nar/gkl560. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 68.Cahuzac B, Berthonneau E, Birlirakis N, Guittet E, Mirande M. A recurrent RNA-binding domain is appended to eukaryotic aminoacyl-tRNA synthetases. EMBO J. 2000;19(3):445–452. doi: 10.1093/emboj/19.3.445. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 69.Cerini C, Semeriva M, Gratecos D. Evolution of the aminoacyl-tRNA synthetase family and the organization of the Drosophila glutamyl-prolyl-tRNA synthetase gene. Intron/exon structure of the gene, control of expression of the two mRNAs, selective advantage of the multienzyme complex. Eur J Bioch. 1997;244(1):176–185. doi: 10.1111/j.1432-1033.1997.00176.x. [DOI] [PubMed] [Google Scholar]
- 70.Ewalt KL, Yang XL, Otero FJ, Liu J, Slike B, Schimmel P. Variant of human enzyme sequesters reactive intermediate. Biochemistry. 2005;44(11):4216–4221. doi: 10.1021/bi048116l. [DOI] [PubMed] [Google Scholar]
- 71.Jia J, Arif A, Ray PS, Fox PL. WHEP domains direct noncanonical function of glutamyl-prolyl tRNA synthetase in translational control of gene expression. Mol cell. 2008;29(6):679–690. doi: 10.1016/j.molcel.2008.01.010. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 72.Kerjan P, Triconnet M, Waller JP. Mammalian prolyl-tRNA synthetase corresponds to the approximately 150 kDa subunit of the high-M(r) aminoacyl-tRNA synthetase complex. Biochimie. 1992;74(2):195–205. doi: 10.1016/0300-9084(92)90046-h. [DOI] [PubMed] [Google Scholar]
- 73.Mukhopadhyay R, Jia J, Arif A, Ray PS, Fox PL. The GAIT system: a gatekeeper of inflammatory gene expression. Trends Biochem Sci. 2009;34(7):324–331. doi: 10.1016/j.tibs.2009.03.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 74.Rho SB, Lee JS, Jeong EJ, Kim KS, Kim YG, Kim S. A multifunctional repeated motif is present in human bifunctional tRNA synthetase. J Biol Chem. 1998;273(18):11267–11273. doi: 10.1074/jbc.273.18.11267. [DOI] [PubMed] [Google Scholar]
- 75.Yang XL, Otero FJ, Skene RJ, McRee DE, Schimmel P, Ribas de Pouplana L. Crystal structures that suggest late development of genetic code components for differentiating aromatic side chains. Proc Natl Acad Sci U S A. 2003;100(26):15376–15380. doi: 10.1073/pnas.2136794100. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 76.Kaminska M, Deniziak M, Kerjan P, Barciszewski J, Mirande M. A recurrent general RNA binding domain appended to plant methionyl-tRNA synthetase acts as a cis-acting cofactor for aminoacylation. EMBO J. 2000;19(24):6908–6917. doi: 10.1093/emboj/19.24.6908. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 77.Kim Y, Shin J, Li R, Cheong C, Kim K, Kim S. A novel anti-tumor cytokine contains an RNA binding motif present in aminoacyl-tRNA synthetases. J Biol Chem. 2000;275(35):27062–27068. doi: 10.1074/jbc.C000216200. [DOI] [PubMed] [Google Scholar]
- 78.Quevillon S, Agou F, Robinson JC, Mirande M. The p43 component of the mammalian multi-synthetase complex is likely to be the precursor of the endothelial monocyte-activating polypeptide II cytokine. J Biol Chem. 1997;272(51):32573–32579. doi: 10.1074/jbc.272.51.32573. [DOI] [PubMed] [Google Scholar]
- 79.Renault L, Kerjan P, Pasqualato S, Menetrey J, Robinson JC, Kawaguchi S, Vassylyev DG, Yokoyama S, Mirande M, Cherfils J. Structure of the EMAPII domain of human aminoacyl-tRNA synthetase complex reveals evolutionary dimer mimicry. EMBO l. 2001;20(3):570–578. doi: 10.1093/emboj/20.3.570. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 80.Robinson JC, Kerjan P, Mirande M. Macromolecular assemblage of aminoacyl-tRNA synthetases: quantitative analysis of protein-protein interactions and mechanism of complex assembly. J Mol Biol. 2000;304(5):983–994. doi: 10.1006/jmbi.2000.4242. [DOI] [PubMed] [Google Scholar]
- 81.Rose A, Schraegle SJ, Stahlberg EA, Meier I. Coiled-coil protein composition of 22 proteomes--differences and common themes in subcellular infrastructure and traffic control. BMC evolutionary biology. 2005;5:66. doi: 10.1186/1471-2148-5-66. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 82.Vellekamp G, Sihag RK, Deutscher MP. Comparison of the complexed and free forms of rat liver arginyl-tRNA synthetase and origin of the free form. J Biol Chem. 1985;260(17):9843–9847. [PubMed] [Google Scholar]
- 83.Zheng YG, Wei H, Ling C, Xu MG, Wang ED. Two forms of human cytoplasmic arginyl-tRNA synthetase produced from two translation initiations by a single mRNA. Biochemistry. 2006;45(4):1338–1344. doi: 10.1021/bi051675n. [DOI] [PubMed] [Google Scholar]
- 84.Woese CR. On the evolution of the genetic code. Proc Natl Acad Sci U S A. 1965;54(6):1546–1552. doi: 10.1073/pnas.54.6.1546. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 85.Woese CR. On the evolution of cells. Proc Natl Acad Sci U S A. 2002;99(13):8742–8747. doi: 10.1073/pnas.132266999. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 86.Karkhanis VA, Boniecki MT, Poruri K, Martinis SA. A viable amino acid editing activity in the leucyl-tRNA synthetase CP1-splicing domain is not required in the yeast mitochondria. J Biol Chem. 2006;281(44):33217–33225. doi: 10.1074/jbc.M607406200. [DOI] [PubMed] [Google Scholar]
- 87.Karkhanis VA, Mascarenhas AP, Martinis SA. Amino acid toxicities of Escherichia coli that are prevented by leucyl-tRNA synthetase amino acid editing. J Bact. 2007;189(23):8765–8768. doi: 10.1128/JB.01215-07. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 88.Lee JW, Beebe K, Nangle LA, Jang J, Longo-Guess CM, Cook SA, Davisson MT, Sundberg JP, Schimmel P, Ackerman SL. Editing-defective tRNA synthetase causes protein misfolding and neurodegeneration. Nature. 2006;443(7107):50–55. doi: 10.1038/nature05096. [DOI] [PubMed] [Google Scholar]
- 89.Nangle LA, Motta CM, Schimmel P. Global effects of mistranslation from an editing defect in mammalian cells. Chem Biol. 2006;13(10):1091–1100. doi: 10.1016/j.chembiol.2006.08.011. [DOI] [PubMed] [Google Scholar]
- 90.Geslain R, Ribas de Pouplana L. Regulation of RNA function by aminoacylation and editing? Trends Genet. 2004;20(12):604–610. doi: 10.1016/j.tig.2004.09.012. [DOI] [PubMed] [Google Scholar]
- 91.Hendrickson TL, Schimmel P. Transfer RNA-dependent amino acid discrimination by aminoacyl-tRNA synthetases. In: Lapointe J, Brakier-Gingras L, editors. In Translation Mechanisms. Landes Bioscience; Georgetown, TX: 2002. [Google Scholar]
- 92.Pauling L: Festschrift Authur Stöll Siebzigsten Geburtstag. Birkhauser Verlag Basel; Switzerland: 1958. The probability of errors in the process of synthesis of protein molecules; pp. 597–602. [Google Scholar]
- 93.Loftfield RB. The frequency of errors in protein biosynthesis. Biochem J. 1963;89:82–92. doi: 10.1042/bj0890082. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 94.Baldwin AN, Berg P. Transfer ribonucleic acid-induced hydrolysis of valyladenylate bound to isoleucyl ribonucleic acid synthetase. J Biol Chem. 1966;241(4):839–845. [PubMed] [Google Scholar]
- 95.Fersht AR. Editing mechanisms in protein synthesis. Rejection of valine by the isoleucyl-tRNA synthetase. Biochemistry. 1977;16(5):1025–1030. doi: 10.1021/bi00624a034. [DOI] [PubMed] [Google Scholar]
- 96.Lin L, Schimmel P. Mutational analysis suggests the same design for editing activities of two tRNA synthetases. Biochemistry. 1996;35(17):5596–5601. doi: 10.1021/bi960011y. [DOI] [PubMed] [Google Scholar]
- 97.Schmidt E, Schimmel P. Mutational isolation of a sieve for editing in a transfer RNA synthetase. Science. 1994;264(5156):265–267. doi: 10.1126/science.8146659. [DOI] [PubMed] [Google Scholar]
- 98.Schmidt E, Schimmel P. Residues in a class I tRNA synthetase which determine selectivity of amino acid recognition in the context of tRNA. Biochemistry. 1995;34(35):11204–11210. doi: 10.1021/bi00035a028. [DOI] [PubMed] [Google Scholar]
- 99.Nureki O, Vassylyev DG, Tateno M, Shimada A, Nakama T, Fukai S, Konno M, Hendrickson TL, Schimmel P, Yokoyama S. Enzyme structure with two catalytic sites for double-sieve selection of substrate. Science. 1998;280(5363):578–582. doi: 10.1126/science.280.5363.578. [DOI] [PubMed] [Google Scholar]
- 100.Dock-Bregeon A, Sankaranarayanan R, Romby P, Caillet J, Springer M, Rees B, Francklyn CS, Ehresmann C, Moras D. Transfer RNA-mediated editing in threonyl-tRNA synthetase. The class II solution to the double discrimination problem. Cell. 2000;103(6):877–884. doi: 10.1016/s0092-8674(00)00191-4. [DOI] [PubMed] [Google Scholar]
- 101.Sankaranarayanan R, Dock-Bregeon AC, Rees B, Bovee M, Caillet J, Romby P, Francklyn CS, Moras D. Zinc ion mediated amino acid discrimination by threonyl-tRNA synthetase. Nature Struct Biology. 2000;7(6):461–465. doi: 10.1038/75856. [DOI] [PubMed] [Google Scholar]
- 102.An S, Musier-Forsyth K. Trans-editing of Cys-tRNAPro by Haemophilus influenzae YbaK protein. J Biol Chem. 2004;279(41):42359–42362. doi: 10.1074/jbc.C400304200. [DOI] [PubMed] [Google Scholar]
- 103.Chong YE, Yang XL, Schimmel P. Natural homolog of tRNA synthetase editing domain rescues conditional lethality caused by mistranslation. J Biol Chem. 2008;283(44):30073–30078. doi: 10.1074/jbc.M805943200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 104.Hellmann RA, Martinis SA. Defects in transient tRNA translocation bypass tRNA synthetase quality control mechanisms. J Biol Chem. 2009;284(17):11478–11484. doi: 10.1074/jbc.M807395200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 105.Palencia A, Crepin T, Vu MT, Lincecum TL, Jr., Martinis SA, Cusack S. Structural dynamics of the aminoacylation and proofreading functional cycle of bacterial leucyl-tRNA synthetase. Nature Struct Mol Biol. 2012;19(7):677–684. doi: 10.1038/nsmb.2317. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 106.Mascarenhas AP, Martinis SA. Functional segregation of a predicted “hinge” site within the β-strand linkers of Escherichia coli leucyl-tRNA synthetase. Biochemistry. 2008;47(16):4808–4816. doi: 10.1021/bi702494q. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 107.Mascarenhas AP, Martinis SA. A glycine hinge for tRNA-dependent translocation of editing substrates to prevent errors by leucyl-tRNA synthetase. FEBS Lett. 2009;583(21):3443–3447. doi: 10.1016/j.febslet.2009.09.039. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 108.Bishop AC, Nomanbhoy TK, Schimmel P. Blocking site-to-site translocation of a misactivated amino acid by mutation of a class I tRNA synthetase. Proc Natl Acad Sci U S A. 2002;99(2):585–590. doi: 10.1073/pnas.012611299. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 109.Eldred EW, Schimmel PR. Rapid deacylation by isoleucyl transfer ribonucleic acid synthetase of isoleucine-specific transfer ribonucleic acid aminoacylated with valine. J Biol Chem. 1972;247(9):2961–2964. [PubMed] [Google Scholar]
- 110.Jakubowski H. Aminoacyl thioester chemistry of class II aminoacyl-tRNA synthetases. Biochemistry. 1997;36(37):11077–11085. doi: 10.1021/bi970589n. [DOI] [PubMed] [Google Scholar]
- 111.Boniecki MT, Vu MT, Betha AK, Martinis SA. CP1-dependent partitioning of pretransfer and posttransfer editing in leucyl-tRNA synthetase. Proc Natl Acad Sci U S A. 2008;105(49):19223–19228. doi: 10.1073/pnas.0809336105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 112.Dulic M, Cvetesic N, Perona JJ, Gruic-Sovulj I. Partitioning of tRNA-dependent editing between pre- and post-transfer pathways in class I aminoacyl-tRNA synthetases. J Biol Chem. 2010;285(31):23799–23809. doi: 10.1074/jbc.M110.133553. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 113.Gruic-Sovulj I, Uter N, Bullock T, Perona JJ. tRNA-dependent aminoacyl-adenylate hydrolysis by a nonediting class I aminoacyl-tRNA synthetase. J Biol Chem. 2005;280(25):23978–23986. doi: 10.1074/jbc.M414260200. [DOI] [PubMed] [Google Scholar]
- 114.Minajigi A, Francklyn CS. Aminoacyl transfer rate dictates choice of editing pathway in threonyl-tRNA synthetase. J Biol Chem. 2010;285(31):23810–23817. doi: 10.1074/jbc.M110.105320. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 115.Lincecum TL, Jr., Tukalo M, Yaremchuk A, Mursinna RS, Williams AM, Sproat BS, Van Den Eynde W, Link A, Van Calenbergh S, Grøtli M, et al. Structural and mechanistic basis of pre- and post-transfer editing by leucyl-tRNA synthetase. Mol Cell. 2003;11(4):951–963. doi: 10.1016/s1097-2765(03)00098-4. [DOI] [PubMed] [Google Scholar]
- 116.Li L, Boniecki MT, Jaffe JD, Imai BS, Yau PM, Luthey-Schulten ZA, Martinis SA. Naturally occurring aminoacyl-tRNA synthetases editing-domain mutations that cause mistranslation in Mycoplasma parasites. Proc Natl Acad Sci U S A. 2011;108(23):9378–9383. doi: 10.1073/pnas.1016460108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 117.Li L, Palencia A, Lukk T, Li Z, Luthey-Schulten ZA, Cusack S, Martinis SA, Boniecki MT. Leucyl-tRNA synthetase editing domain functions as a molecular rheostat to control codon ambiguity in Mycoplasma pathogens. Proc Natl Acad Sci U S A. 2013;110(10):3817–3822. doi: 10.1073/pnas.1218374110. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 118.Englisch S, Englisch U, von der Haar F, Cramer F. The proofreading of hydroxy analogues of leucine and isoleucine by leucyl-tRNA synthetases from E. coli and yeast. Nucl Acids Res. 1986;14(19):7529–7539. doi: 10.1093/nar/14.19.7529. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 119.Fersht AR, Dingwall C. Evidence for the double-sieve editing mechanism in protein synthesis. Steric exclusion of isoleucine by valyl-tRNA synthetases. Biochemistry. 1979;18(12):2627–2631. doi: 10.1021/bi00579a030. [DOI] [PubMed] [Google Scholar]
- 120.Martinis SA, Boniecki MT. The balance between pre- and post-transfer editing in tRNA synthetases. FEBS Lett. 2010;584(2):455–459. doi: 10.1016/j.febslet.2009.11.071. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 121.Williams AM, Martinis SA. Mutational unmasking of a tRNA-dependent pathway for preventing genetic code ambiguity. Proc Natl Acad Sci U S A. 2006;103(10):3586–3591. doi: 10.1073/pnas.0507362103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 122.Splan KE, Musier-Forsyth K, Boniecki MT, Martinis SA. In vitro assays for the determination of aminoacyl-tRNA synthetase editing activity. Methods. 2008;44(2):119–128. doi: 10.1016/j.ymeth.2007.10.009. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 123.Chen X, Ma JJ, Tan M, Yao P, Hu QH, Eriani G, Wang ED. Modular pathways for editing non-cognate amino acids by human cytoplasmic leucyl-tRNA synthetase. Nucl Acids Res. 2011;39(1):235–247. doi: 10.1093/nar/gkq763. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 124.Sarkar J, Martinis SA. Amino-acid-dependent shift in tRNA synthetase editing mechanisms. J Am Chem Soc. 2011;133(46):18510–18513. doi: 10.1021/ja2048122. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 125.Pang YL, Martinis SA. A paradigm shift for the amino acid editing mechanism of human cytoplasmic leucyl-tRNA synthetase. Biochemistry. 2009;48(38):8958–8964. doi: 10.1021/bi901111y. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 126.Santos MAS, Tuite MF. The CUG codon is decoded in vivo as serine and not leucine in Candida albicans. Nucl Acids Res. 1995;23(9):1481–1486. doi: 10.1093/nar/23.9.1481. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 127.Gomes AC, Miranda I, Silva RM, Moura GR, Thomas B, Akoulitchev A, Santos MA. A genetic code alteration generates a proteome of high diversity in the human pathogen Candida albicans. Genome Biol. 2007;8(10):R206. doi: 10.1186/gb-2007-8-10-r206. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 128.Kim S-H. Crystal structure of yeast tRNAPhe and general structural features of other tRNAs. In: Schimmel P, Söll D, Abelson JM, editors. Transfer RNA: Structure, Properties, and Recognition. Cold Spring Harbor Laboratory; Cold Spring Harbor, NY: 1979. pp. 83–100. [Google Scholar]
- 129.Hou YM, Schimmel P. A simple structural feature is a major determinant of the identity of a transfer RNA. Nature. 1988;333(6169):140–145. doi: 10.1038/333140a0. [DOI] [PubMed] [Google Scholar]
- 130.McClain WH, Foss K. Changing the identity of a tRNA by introducing a G-U wobble pair near the 3′ acceptor end. Science. 1988;240(4853):793–796. doi: 10.1126/science.2452483. [DOI] [PubMed] [Google Scholar]
- 131.Normanly J, Ollick T, Abelson J. Eight base changes are sufficient to convert a leucine-inserting tRNA into a serine-inserting tRNA. Proc Natl Acad Sci U S A. 1992;89(12):5680–5684. doi: 10.1073/pnas.89.12.5680. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 132.Larkin DC, Williams AM, Martinis SA, Fox GE. Identification of essential domains for Escherichia coli tRNALeu aminoacylation and amino acid editing using minimalist RNA molecules. Nucl Acids Res. 2002;30(10):2103–2113. doi: 10.1093/nar/30.10.2103. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 133.Crothers DM, Seno T, Söll G. Is there a discriminator site in transfer RNA? Proc Natl Acad Sci U S A. 1972;69(10):3063–3067. doi: 10.1073/pnas.69.10.3063. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 134.McClain WH, Chen YM, Foss K, Schneider J. Association of transfer RNA acceptor identity with a helical irregularity. Science. 1988;242(4886):1681–1684. doi: 10.1126/science.2462282. [DOI] [PubMed] [Google Scholar]
- 135.Achsel T, Gross HJ. Identity determinants of human tRNASer: sequence elements necessary for serylation and maturation of a tRNA with a long extra arm. EMBO J. 1993;12(8):3333–3338. doi: 10.1002/j.1460-2075.1993.tb06003.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 136.Muramatsu T, Nishikawa K, Nemoto F, Kuchino Y, Nishimura S, Miyazawa T, Yokoyama S. Codon and amino-acid specificities of a transfer RNA are both converted by a single post-transcriptional modification. Nature. 1988;336(6195):179–181. doi: 10.1038/336179a0. [DOI] [PubMed] [Google Scholar]
- 137.Perret V, Garcia A, Grosjean H, Ebel JP, Florentz C, Giegé R. Relaxation of a transfer RNA specificity by removal of modified nucleotides. Nature. 1990;344(6268):787–789. doi: 10.1038/344787a0. [DOI] [PubMed] [Google Scholar]
- 138.Putz J, Florentz C, Benseler F, Giegé R. A single methyl group prevents the mischarging of a tRNA. Nature Struct Biol. 1994;1(9):580–582. doi: 10.1038/nsb0994-580. [DOI] [PubMed] [Google Scholar]
- 139.Breitschopf K, Achsel T, Busch K, Gross HJ. Identity elements of human tRNALeu: structural requirements for converting human tRNASer into a leucine acceptor in vitro. Nucl Acids Res. 1995;23(18):3633–3637. doi: 10.1093/nar/23.18.3633. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 140.Soma A, Kumagai R, Nishikawa K, Himeno H. The anticodon loop is a major identity determinant of Saccharomyces cerevisiae tRNALeu. J Mol Biol. 1996;263(5):707–714. doi: 10.1006/jmbi.1996.0610. [DOI] [PubMed] [Google Scholar]
- 141.Sethi A, Eargle J, Black AA, Luthey-Schulten Z. Dynamical networks in tRNA:protein complexes. Proc Natl Acad Sci U S A. 2009;106(16):6620–6625. doi: 10.1073/pnas.0810961106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 142.Lee SW, Cho BH, Park SG, Kim S. Aminoacyl-tRNA synthetase complexes: beyond translation. J Cell Sci. 2004;117(Pt 17):3725–3734. doi: 10.1242/jcs.01342. [DOI] [PubMed] [Google Scholar]
- 143.Deutscher MP. The eucaryotic aminoacyl-tRNA synthetase complex: Suggestions for its structure and function. J. Cell Biol. 1984;99:373–377. doi: 10.1083/jcb.99.2.373. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 144.Ray PS, Arif A, Fox PL. Macromolecular complexes as depots for releasable regulatory proteins. Trends Biochem Sci. 2007;32(4):158–164. doi: 10.1016/j.tibs.2007.02.003. [DOI] [PubMed] [Google Scholar]
- 145.Berthonneau E, Mirande M. A gene fusion event in the evolution of aminoacyl-tRNA synthetases. FEBS Letté. 2000;470(3):300–304. doi: 10.1016/s0014-5793(00)01343-0. [DOI] [PubMed] [Google Scholar]
- 146.Jeong EJ, Hwang GS, Kim KH, Kim MJ, Kim S, Kim KS. Structural analysis of multifunctional peptide motifs in human bifunctional tRNA synthetase: identification of RNA-binding residues and functional implications for tandem repeats. Biochemistry. 2000;39(51):15775–15782. doi: 10.1021/bi001393h. [DOI] [PubMed] [Google Scholar]
- 147.Kaminska M, Havrylenko S, Decottignies P, Gillet S, Le Marechal P, Negrutskii B, Mirande M. Dissection of the structural organization of the aminoacyl-tRNA synthetase complex. J Biol Chem. 2009;284(10):6053–6060. doi: 10.1074/jbc.M809636200. [DOI] [PubMed] [Google Scholar]
- 148.Park SG, Ewalt KL, Kim S. Functional expansion of aminoacyl-tRNA synthetases and their interacting factors: new perspectives on housekeepers. Trends Biochem Sci. 2005;30(10):569–574. doi: 10.1016/j.tibs.2005.08.004. [DOI] [PubMed] [Google Scholar]
- 149.Norcum MT, Boisset N. Three-dimensional architecture of the eukaryotic multisynthetase complex determined from negatively stained and cryoelectron micrographs. FEBS Lett. 2002;512(1-3):298–302. doi: 10.1016/s0014-5793(02)02262-7. [DOI] [PubMed] [Google Scholar]
- 150.Norcum MT, Warrington JA. Structural analysis of the multienzyme aminoacyl-tRNA synthetase complex: a three-domain model based on reversible chemical crosslinking. Protein Sci. 1998;7(1):79–87. doi: 10.1002/pro.5560070108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 151.Ahn HC, Kim S, Lee BJ. Solution structure and p43 binding of the p38 leucine zipper motif: coiled-coil interactions mediate the association between p38 and p43. FEBS Lett. 2003;542(1-3):119–124. doi: 10.1016/s0014-5793(03)00362-4. [DOI] [PubMed] [Google Scholar]
- 152.Norcum MT, Warrington JA. The cytokine portion of p43 occupies a central position within the eukaryotic multisynthetase complex. J Biol Chem. 2000;275(24):17921–17924. doi: 10.1074/jbc.C000266200. [DOI] [PubMed] [Google Scholar]
- 153.Park SG, Jung KH, Lee JS, Jo YJ, Motegi H, Kim S, Shiba K. Precursor of pro-apoptotic cytokine modulates aminoacylation activity of tRNA synthetase. J Biol Chem. 1999;274(24):16673–16676. doi: 10.1074/jbc.274.24.16673. [DOI] [PubMed] [Google Scholar]
- 154.Yannay-Cohen N, Carmi-Levy I, Kay G, Yang CM, Han JM, Kemeny DM, Kim S, Nechushtan H, Razin E. LysRS serves as a key signaling molecule in the immune response by regulating gene expression. Mol Cell. 2009;34(5):603–611. doi: 10.1016/j.molcel.2009.05.019. [DOI] [PubMed] [Google Scholar]
- 155.Han JM, Park BJ, Park SG, Oh YS, Choi SJ, Lee SW, Hwang SK, Chang SH, Cho MH, Kim S. AIMP2/p38, the scaffold for the multi-tRNA synthetase complex, responds to genotoxic stresses via p53. Proc Natl Acad Sci U S A. 2008;105(32):11206–11211. doi: 10.1073/pnas.0800297105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 156.Han JM, Kwon NH, Lee JY, Jeong SJ, Jung HJ, Kim HR, Li Z, Kim S. Identification of gp96 as a novel target for treatment of autoimmune disease in mice. PloS One. 2010;5(3):e9792. doi: 10.1371/journal.pone.0009792. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 157.Kim G, Han JM, Kim S. Toll-like receptor 4-mediated c-Jun N-terminal kinase activation induces gp96 cell surface expression via AIMP1 phosphorylation. Biochem Biophys Res Comm. 2010;397(1):100–105. doi: 10.1016/j.bbrc.2010.05.075. [DOI] [PubMed] [Google Scholar]
- 158.Lee YN, Nechushtan H, Figov N, Razin E. The function of lysyl-tRNA synthetase and Ap4A as signaling regulators of MITF activity in FcepsilonRI-activated mast cells. Immunity. 2004;20(2):145–151. doi: 10.1016/s1074-7613(04)00020-2. [DOI] [PubMed] [Google Scholar]
- 159.Park SG, Kim HJ, Min YH, Choi EC, Shin YK, Park BJ, Lee SW, Kim S. Human lysyl-tRNA synthetase is secreted to trigger proinflammatory response. Proc Natl Acad Sci U S A. 2005;102(18):6356–6361. doi: 10.1073/pnas.0500226102. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 160.Nechushtan H, Kim S, Kay G, Razin E. The physiological role of lysyl tRNA synthetase in the immune system. Adv Immunology. 2009;103:1–27. doi: 10.1016/S0065-2776(09)03001-6. [DOI] [PubMed] [Google Scholar]
- 161.Mazumder B, Sampath P, Seshadri V, Maitra RK, DiCorleto PE, Fox PL. Regulated release of L13a from the 60S ribosomal subunit as a mechanism of transcript-specific translational control. Cell. 2003;115(2):187–198. doi: 10.1016/s0092-8674(03)00773-6. [DOI] [PubMed] [Google Scholar]
- 162.Sampath P, Mazumder B, Seshadri V, Gerber CA, Chavatte L, Kinter M, Ting SM, Dignam JD, Kim S, Driscoll DM, et al. Noncanonical function of glutamyl-prolyl-tRNA synthetase: gene-specific silencing of translation. Cell. 2004;119(2):195–208. doi: 10.1016/j.cell.2004.09.030. [DOI] [PubMed] [Google Scholar]
- 163.Simos G, Segref A, Fasiolo F, Hellmuth K, Shevchenko A, Mann M, Hurt EC. The yeast protein Arc1p binds to tRNA and functions as a cofactor for the methionyl- and glutamyl-tRNA synthetases. EMBO J. 1996;15(19):5437–5448. [PMC free article] [PubMed] [Google Scholar]
- 164.Frechin M, Senger B, Brayé M, Kern D, Martin RP, Becker HD. Yeast mitochondrial Gln-tRNA(Gln) is generated by a GatFAB-mediated transamidation pathway involving Arc1p-controlled subcellular sorting of cytosolic GluRS. Genes Dev. 2009;23(9):1119–1130. doi: 10.1101/gad.518109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 165.Praetorius-Ibba M, Hausmann CD, Paras M, Rogers TE, Ibba M. Functional association between three archaeal aminoacyl-tRNA synthetases. J Biol Chem. 2007;282(6):3680–3687. doi: 10.1074/jbc.M609988200. [DOI] [PubMed] [Google Scholar]
- 166.Praetorius-Ibba M, Rogers TE, Samson R, Kelman Z, Ibba M. Association between Archaeal prolyl- and leucyl-tRNA synthetases enhances tRNAPro aminoacylation. J Biol Chem. 2005;280(28):26099–26104. doi: 10.1074/jbc.M503539200. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 167.Hausmann CD, Praetorius-Ibba M, Ibba M. An aminoacyl-tRNA synthetase:elongation factor complex for substrate channeling in archaeal translation. Nucl Acids Res. 2007;35(18):6094–6102. doi: 10.1093/nar/gkm534. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 168.Park SG, Shin H, Shin YK, Lee Y, Choi EC, Park BJ, Kim S. The novel cytokine p43 stimulates dermal fibroblast proliferation and wound repair. Am J Pathol. 2005;166(2):387–398. doi: 10.1016/S0002-9440(10)62262-6. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 169.Han JM, Park SG, Lee Y, Kim S. Structural separation of different extracellular activities in aminoacyl-tRNA synthetase-interacting multi-functional protein, p43/AIMP1. Biochem Biophys Res Comm. 2006;342(1):113–118. doi: 10.1016/j.bbrc.2006.01.117. [DOI] [PubMed] [Google Scholar]
- 170.Park SG, Kang YS, Ahn YH, Lee SH, Kim KR, Kim KW, Koh GY, Ko YG, Kim S. Dose-dependent biphasic activity of tRNA synthetase-associating factor, p43, in angiogenesis. J Biol Chem. 2002;277(47):45243–45248. doi: 10.1074/jbc.M207934200. [DOI] [PubMed] [Google Scholar]
- 171.Han JM, Park SG, Liu B, Park BJ, Kim JY, Jin CH, Song YW, Li Z, Kim S. Aminoacyl-tRNA synthetase-interacting multifunctional protein 1/p43 controls endoplasmic reticulum retention of heat shock protein gp96: its pathological implications in lupus-like autoimmune diseases. Am J Pathol. 2007;170(6):2042–2054. doi: 10.2353/ajpath.2007.061266. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 172.Ko HS, von Coelln R, Sriram SR, Kim SW, Chung KK, Pletnikova O, Troncoso J, Johnson B, Saffary R, Goh EL, et al. Accumulation of the authentic parkin substrate aminoacyl-tRNA synthetase cofactor, p38/JTV-1, leads to catecholaminergic cell death. J Neurosci. 2005;25(35):7968–7978. doi: 10.1523/JNEUROSCI.2172-05.2005. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 173.Park BJ, Kang JW, Lee SW, Choi SJ, Shin YK, Ahn YH, Choi YH, Choi D, Lee KS, Kim S. The haploinsufficient tumor suppressor p18 upregulates p53 via interactions with ATM/ATR. Cell. 2005;120(2):209–221. doi: 10.1016/j.cell.2004.11.054. [DOI] [PubMed] [Google Scholar]
- 174.Martinis SA, Plateau P, Cavarelli J, Florentz C. Aminoacyl-tRNA synthetases: a family of expanding functions. EMBO J. 1999;18(17):4591–4596. doi: 10.1093/emboj/18.17.4591. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 175.Park SG, Schimmel P, Kim S. Aminoacyl tRNA synthetases and their connections to disease. Proc Natl Acad Sci U S A. 2008;105(32):11043–11049. doi: 10.1073/pnas.0802862105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 176.Herbert CJ, Labouesse M, Dujardin G, Slonimski PP. The NAM2 proteins from S. cerevisiae and S. douglasii are mitochondrial leucyl-tRNA synthetases, and are involved in mRNA splicing. EMBO J. 1988;7(2):473–483. doi: 10.1002/j.1460-2075.1988.tb02835.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 177.Kämper U, Kuck U, Cherniack AD, Lambowitz AM. The mitochondrial tyrosyl-tRNA synthetase of Podospora anserina is a bifunctional enzyme active in protein synthesis and RNA splicing. Mol Cell Biol. 1992;12(2):499–511. doi: 10.1128/mcb.12.2.499. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 178.Labouesse M. The yeast mitochondrial leucyl-tRNA synthetase is a splicing factor for the excision of several group I introns. Mol Gen Genet. 1990;224(2):209–221. doi: 10.1007/BF00271554. [DOI] [PubMed] [Google Scholar]
- 179.Labouesse M, Dujardin G, Slonimski PP. The yeast nuclear gene NAM2 is essential for mitochondrial DNA integrity and can cure a mitochondrial RNA-maturase deficiency. Cell. 1985;41(1):133–143. doi: 10.1016/0092-8674(85)90068-6. [DOI] [PubMed] [Google Scholar]
- 180.Romby P, Caillet J, Ebel C, Sacerdot C, Graffe M, Eyermann F, Brunel C, Moine H, Ehresmann C, Ehresmann B, et al. The expression of E. coli threonyl-tRNA synthetase is regulated at the translational level by symmetrical operator-repressor interactions. EMBO J. 1996;15(21):5976–5987. [PMC free article] [PubMed] [Google Scholar]
- 181.Condon C, Grunberg-Manago M, Putzer H. Aminoacyl-tRNA synthetase gene regulation in Bacillus subtilis. Biochimie. 1996;78(6):381–389. doi: 10.1016/0300-9084(96)84744-4. [DOI] [PubMed] [Google Scholar]
- 182.Putney SD, Schimmel P. An aminoacyl tRNA synthetase binds to a specific DNA sequence and regulates its gene transcription. Nature. 1981;291(5817):632–635. doi: 10.1038/291632a0. [DOI] [PubMed] [Google Scholar]
- 183.Martinis SA, Joy Pang YL. Jekyll & Hyde: evolution of a superfamily. Chem Biol. 2007;14(12):1307–1308. doi: 10.1016/j.chembiol.2007.12.001. [DOI] [PubMed] [Google Scholar]
- 184.Antonellis A, Ellsworth RE, Sambuughin N, Puls I, Abel A, Lee-Lin SQ, Jordanova A, Kremensky I, Christodoulou K, Middleton LT, et al. Glycyl tRNA synthetase mutations in Charcot-Marie-Tooth disease type 2D and distal spinal muscular atrophy type V. Am J Hum Gen. 2003;72(5):1293–1299. doi: 10.1086/375039. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 185.Jordanova A, Irobi J, Thomas FP, Van Dijck P, Meerschaert K, Dewil M, Dierick I, Jacobs A, De Vriendt E, Guergueltcheva V, et al. Disrupted function and axonal distribution of mutant tyrosyl-tRNA synthetase in dominant intermediate Charcot-Marie-Tooth neuropathy. Nature Gen. 2006;38(2):197–202. doi: 10.1038/ng1727. [DOI] [PubMed] [Google Scholar]
- 186.Storkebaum E, Leitao-Goncalves R, Godenschwege T, Nangle L, Mejia M, Bosmans I, Ooms T, Jacobs A, Van Dijck P, Yang XL, et al. Dominant mutations in the tyrosyl-tRNA synthetase gene recapitulate in Drosophila features of human Charcot-Marie-Tooth neuropathy. Proc Natl Acad Sci U S A. 2009;106(28):11782–11787. doi: 10.1073/pnas.0905339106. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 187.Nangle LA, Zhang W, Xie W, Yang XL, Schimmel P. Charcot-Marie-Tooth disease-associated mutant tRNA synthetases linked to altered dimer interface and neurite distribution defect. Proc Natl Acad Sci U S A. 2007;104(27):11239–11244. doi: 10.1073/pnas.0705055104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 188.Mathews MB, Bernstein RM. Myositis autoantibody inhibits histidyl-tRNA synthetase: a model for autoimmunity. Nature. 1983;304(5922):177–179. doi: 10.1038/304177a0. [DOI] [PubMed] [Google Scholar]
- 189.Raben N, Nichols R, Dohlman J, McPhie P, Sridhar V, Hyde C, Leff R, Plotz P. A motif in human histidyl-tRNA synthetase which is shared among several aminoacyl-tRNA synthetases is a coiled-coil that is essential for enzymatic activity and contains the major autoantigenic epitope. J Biol Chem. 1994;269(39):24277–24283. [PubMed] [Google Scholar]
- 190.Beaulande M, Tarbouriech N, Hartlein M. Human cytosolic asparaginyl-tRNA synthetase: cDNA sequence, functional expression in Escherichia coli and characterization as human autoantigen. Nucl Acids Res. 1998;26(2):521–524. doi: 10.1093/nar/26.2.521. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 191.Hirakata M, Suwa A, Nagai S, Kron MA, Trieu EP, Mimori T, Akizuki M, Targoff IN. Anti-KS: identification of autoantibodies to asparaginyl-transfer RNA synthetase associated with interstitial lung disease. J Immunol. 1999;162(4):2315–2320. [PubMed] [Google Scholar]
- 192.Hirakata M, Suwa A, Takeda Y, Matsuoka Y, Irimajiri S, Targoff IN, Hardin JA, Craft J. Autoantibodies to glycyl-transfer RNA synthetase in myositis. Association with dermatomyositis and immunologic heterogeneity. Arthritis and Rheumatism. 1996;39(1):146–151. doi: 10.1002/art.1780390119. [DOI] [PubMed] [Google Scholar]
- 193.Targoff IN. Autoantibodies to aminoacyl-transfer RNA synthetases for isoleucine and glycine. Two additional synthetases are antigenic in myositis. J Immunol. 1990;144(5):1737–1743. [PubMed] [Google Scholar]
- 194.Targoff IN, Trieu EP, Miller FW. Reaction of anti-OJ autoantibodies with components of the multi-enzyme complex of aminoacyl-tRNA synthetases in addition to isoleucyl-tRNA synthetase. J Clin Inv. 1993;91(6):2556–2564. doi: 10.1172/JCI116493. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 195.Castro de Moura M, Miro F, Han JM, Kim S, Celada A, Ribas de Pouplana L. Entamoeba lysyl-tRNA synthetase contains a cytokine-like domain with chemokine activity towards human endothelial cells. PLoS Neglected Tropical Diseases. 2011;5(11):e1398. doi: 10.1371/journal.pntd.0001398. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 196.Sajish M, Zhou Q, Kishi S, Valdez DM, Jr., Kapoor M, Guo M, Lee S, Kim S, Yang XL, Schimmel P. Trp-tRNA synthetase bridges DNA-PKcs to PARP-1 to link IFN-gamma and p53 signaling. Nature Chem Biol. 2012;8(6):547–554. doi: 10.1038/nchembio.937. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 197.Park MC, Kang T, Jin D, Han JM, Kim SB, Park YJ, Cho K, Park YW, Guo M, He W, et al. Secreted human glycyl-tRNA synthetase implicated in defense against ERK-activated tumorigenesis. Proc Natl Acad Sci U S A. 2012;109(11):E640–647. doi: 10.1073/pnas.1200194109. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 198.Howard OM, Dong HF, Yang D, Raben N, Nagaraju K, Rosen A, Casciola-Rosen L, Hartlein M, Kron M, Yang D, et al. Histidyl-tRNA synthetase and asparaginyl-tRNA synthetase, autoantigens in myositis, activate chemokine receptors on T lymphocytes and immature dendritic cells. J Exp Med. 2002;196(6):781–791. doi: 10.1084/jem.20020186. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 199.Kepp O, Gdoura A, Martins I, Panaretakis T, Schlemmer F, Tesniere A, Fimia GM, Ciccosanti F, Burgevin A, Piacentini M, et al. Lysyl tRNA synthetase is required for the translocation of calreticulin to the cell surface in immunogenic death. Cell Cycle. 2010;9(15):3072–3077. doi: 10.4161/cc.9.15.12459. [DOI] [PubMed] [Google Scholar]
- 200.Arif A, Jia J, Moodt RA, DiCorleto PE, Fox PL. Phosphorylation of glutamyl-prolyl tRNA synthetase by cyclin-dependent kinase 5 dictates transcript-selective translational control. Proc Natl Acad Sci U S A. 2011;108(4):1415–1420. doi: 10.1073/pnas.1011275108. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 201.Yao P, Potdar AA, Arif A, Ray PS, Mukhopadhyay R, Willard B, Xu Y, Yan J, Saidel GM, Fox PL. Coding region polyadenylation generates a truncated tRNA synthetase that counters translation repression. Cell. 2012;149(1):88–100. doi: 10.1016/j.cell.2012.02.018. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 202.Park SJ, Kim SH, Choi HS, Rhee Y, Lim SK. Fibroblast growth factor 2-induced cytoplasmic asparaginyl-tRNA synthetase promotes survival of osteoblasts by regulating anti-apoptotic PI3K/Akt signaling. Bone. 2009;45(5):994–1003. doi: 10.1016/j.bone.2009.07.018. [DOI] [PubMed] [Google Scholar]
- 203.Ramirez BL, Howard OM, Dong HF, Edamatsu T, Gao P, Härtlein M, Kron M. Brugia malayi asparaginyl-transfer RNA synthetase induces chemotaxis of human leukocytes and activates G-protein-coupled receptors CXCR1 and CXCR2. The J Infect Dis. 2006;193(8):1164–1171. doi: 10.1086/501369. [DOI] [PubMed] [Google Scholar]
- 204.Houman F, Rho SB, Zhang J, Shen X, Wang CC, Schimmel P, Martinis SA. A prokaryote and human tRNA synthetase provide an essential RNA splicing function in yeast mitochondria. Proc Natl Acad Sci U S A. 2000;97(25):13743–13748. doi: 10.1073/pnas.240465597. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 205.Rho SB, Lincecum TL, Jr., Martinis SA. An inserted region of leucyl-tRNA synthetase plays a critical role in group I intron splicing. EMBO J. 2002;21(24):6874–6881. doi: 10.1093/emboj/cdf671. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 206.Boniecki MT, Rho SB, Tukalo M, Hsu JL, Romero EP, Martinis SA. Leucyl-tRNA synthetase-dependent and -independent activation of a group I intron. J Biol Chem. 2009;284(39):26243–26250. doi: 10.1074/jbc.M109.031179. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 207.Caprara MG, Mohr G, Lambowitz AM. A tyrosyl-tRNA synthetase protein induces tertiary folding of the group I intron catalytic core. J Mol Biol. 1996;257(3):512–531. doi: 10.1006/jmbi.1996.0182. [DOI] [PubMed] [Google Scholar]
- 208.Cen S, Khorchid A, Javanbakht H, Gabor J, Stello T, Shiba K, Musier-Forsyth K, Kleiman L. Incorporation of lysyl-tRNA synthetase into human immunodeficiency virus type 1. J Virology. 2001;75(11):5043–5048. doi: 10.1128/JVI.75.11.5043-5048.2001. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 209.Kleiman L, Jones CP, Musier-Forsyth K. Formation of the tRNALys packaging complex in HIV-1. FEBS Lett. 2010;584(2):359–365. doi: 10.1016/j.febslet.2009.11.038. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 210.Halwani R, Cen S, Javanbakht H, Saadatmand J, Kim S, Shiba K, Kleiman L. Cellular distribution of Lysyl-tRNA synthetase and its interaction with Gag during human immunodeficiency virus type 1 assembly. J Virol. 2004;78(14):7553–7564. doi: 10.1128/JVI.78.14.7553-7564.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 211.Javanbakht H, Halwani R, Cen S, Saadatmand J, Musier-Forsyth K, Gottlinger H, Kleiman L. The interaction between HIV-1 Gag and human lysyl-tRNA synthetase during viral assembly. J Biol Chem. 2003;278(30):27644–27651. doi: 10.1074/jbc.M301840200. [DOI] [PubMed] [Google Scholar]
- 212.Cen S, Javanbakht H, Niu M, Kleiman L. Ability of wild-type and mutant lysyl-tRNA synthetase to facilitate tRNALys incorporation into human immunodeficiency virus type 1. J Virology. 2004;78(3):1595–1601. doi: 10.1128/JVI.78.3.1595-1601.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 213.Gabor J, Cen S, Javanbakht H, Niu M, Kleiman L. Effect of altering the tRNALys3 concentration in human immunodeficiency virus type 1 upon its annealing to viral RNA, GagPol incorporation, and viral infectivity. J Virology. 2002;76(18):9096–9102. doi: 10.1128/JVI.76.18.9096-9102.2002. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 214.Javanbakht H, Cen S, Musier-Forsyth K, Kleiman L. Correlation between tRNA Lys3 aminoacylation and its incorporation into HIV-1. J Biol Chem. 2002;277(20):17389–17396. doi: 10.1074/jbc.M112479200. [DOI] [PubMed] [Google Scholar]
- 215.Kobbi L, Octobre G, Dias J, Comisso M, Mirande M. Association of mitochondrial Lysyl-tRNA synthetase with HIV-1 GagPol involves catalytic domain of the synthetase and transframe and integrase domains of Pol. J Mol Biol. 2011;410(5):875–886. doi: 10.1016/j.jmb.2011.03.005. [DOI] [PubMed] [Google Scholar]
- 216.Jiang M, Mak J, Ladha A, Cohen E, Klein M, Rovinski B, Kleiman L. Identification of tRNAs incorporated into wild-type and mutant human immunodeficiency virus type 1. Journal of virology. 1993;67(6):3246–3253. doi: 10.1128/jvi.67.6.3246-3253.1993. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 217.Keiler KC, Waller PR, Sauer RT. Role of a peptide tagging system in degradation of proteins synthesized from damaged messenger RNA. Science. 1996;271(5251):990–993. doi: 10.1126/science.271.5251.990. [DOI] [PubMed] [Google Scholar]
- 218.Stewart TS, Roberts RJ, Strominger JL. Novel species of tRNA. Nature. 1971;230(5288):36–38. doi: 10.1038/230036a0. [DOI] [PubMed] [Google Scholar]
- 219.Roy H, Ibba M. RNA-dependent lipid remodeling by bacterial multiple peptide resistance factors. Proc Natl Acad Sci U S A. 2008;105(12):4667–4672. doi: 10.1073/pnas.0800006105. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 220.Dreher TW. Viral tRNAs and tRNA-like structures. Wiley Interdisciplinary Reviews. 2010;1(3):402–414. doi: 10.1002/wrna.42. [DOI] [PubMed] [Google Scholar]
- 221.Dreher TW, Tsai CH, Florentz C, Giegé R. Specific valylation of turnip yellow mosaic virus RNA by wheat germ valyl-tRNA synthetase determined by three anticodon loop nucleotides. Biochemistry. 1992;31(38):9183–9189. doi: 10.1021/bi00153a010. [DOI] [PubMed] [Google Scholar]
- 222.Florentz C, Dreher TW, Rudinger J, Giegé R. Specific valylation identity of turnip yellow mosaic virus RNA by yeast valyl-tRNA synthetase is directed by the anticodon in a kinetic rather than affinity-based discrimination. Eur J Bioch. 1991;195(1):229–234. doi: 10.1111/j.1432-1033.1991.tb15698.x. [DOI] [PubMed] [Google Scholar]
- 223.Felden B, Florentz C, McPherson A, Giegé R. A histidine accepting tRNA-like fold at the 3′-end of satellite tobacco mosaic virus RNA. Nucl Acids Res. 1994;22(15):2882–2886. doi: 10.1093/nar/22.15.2882. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 224.Rietveld K, Linschooten K, Pleij CW, Bosch L. The three-dimensional folding of the tRNA-like structure of tobacco mosaic virus RNA. A new building principle applied twice. EMBO J. 1984;3(11):2613–2619. doi: 10.1002/j.1460-2075.1984.tb02182.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 225.Rudinger J, Felden B, Florentz C, Giegé R. Strategy for RNA recognition by yeast histidyl-tRNA synthetase. Bioorg Med Chem. 1997;5(6):1001–1009. doi: 10.1016/s0968-0896(97)00061-8. [DOI] [PubMed] [Google Scholar]
- 226.Fechter P, Giegé R, Rudinger-Thirion J. Specific tyrosylation of the bulky tRNA-like structure of brome mosaic virus RNA relies solely on identity nucleotides present in its amino acid-accepting domain. J Mol Biol. 2001;309(2):387–399. doi: 10.1006/jmbi.2001.4654. [DOI] [PubMed] [Google Scholar]
- 227.Perret V, Florentz C, Dreher T, Giegé R. Structural analogies between the 3′ tRNA-like structure of brome mosaic virus RNA and yeast tRNATyr revealed by protection studies with yeast tyrosyl-tRNA synthetase. Eur J Biochem. 1989;185(2):331–339. doi: 10.1111/j.1432-1033.1989.tb15120.x. [DOI] [PubMed] [Google Scholar]
- 228.Florentz C, Giegé R. tRNA-Like structures in plant viral RNAs. In: Dieter Söll Uttam RajBhandary, editor. tRNA: Structure, Biosynthesis and Function. ASM Press; Washington, D.C: 1995. [Google Scholar]
- 229.Pleij CW, Rietveld K, Bosch L. A new principle of RNA folding based on pseudoknotting. Nucl Acids Res. 1985;13(5):1717–1731. doi: 10.1093/nar/13.5.1717. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 230.Rietveld K, Pleij CW, Bosch L. Three-dimensional models of the tRNA-like 3′ termini of some plant viral RNAs. EMBOJ. 1983;2(7):1079–1085. doi: 10.1002/j.1460-2075.1983.tb01549.x. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 231.Florentz C, Giegé R. Contact areas of the turnip yellow mosaic virus tRNA-like structure interacting with yeast valyl-tRNA synthetase. J Mol Biol. 1986;191(1):117–130. doi: 10.1016/0022-2836(86)90427-4. [DOI] [PubMed] [Google Scholar]
- 232.Salazar JC, Ambrogelly A, Crain PF, McCloskey JA, Soll D. A truncated aminoacyl-tRNA synthetase modifies RNA. Proc Natl Acad Sci U S A. 2004;101(20):7536–7541. doi: 10.1073/pnas.0401982101. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 233.Campanacci V, Dubois DY, Becker HD, Kern D, Spinelli S, Valencia C, Pagot F, Salomoni A, Grisel S, Vincentelli R, et al. The Escherichia coli YadB gene product reveals a novel aminoacyl-tRNA synthetase like activity. J Mol Biol. 2004;337(2):273–283. doi: 10.1016/j.jmb.2004.01.027. [DOI] [PubMed] [Google Scholar]
- 234.Sissler M, Delorme C, Bond J, Ehrlich SD, Renault P, Francklyn C. An aminoacyl-tRNA synthetase paralog with a catalytic role in histidine biosynthesis. Proc Natl Acad Sci U S A. 1999;96(16):8985–8990. doi: 10.1073/pnas.96.16.8985. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 235.Sutherland R, Boon RJ, Griffin KE, Masters PJ, Slocombe B, White AR. Antibacterial activity of mupirocin (pseudomonic acid), a new antibiotic for topical use. Antimicrob Agents Chemother. 1985;27(4):495–498. doi: 10.1128/aac.27.4.495. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 236.Hughes J, Mellows G. Inhibition of isoleucyl-transfer ribonucleic acid synthetase in Escherichia coli by pseudomonic acid. Biochem J. 1978;176(1):305–318. doi: 10.1042/bj1760305. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 237.Hughes J, Mellows G. Interaction of pseudomonic acid A with Escherichia coli B isoleucyl-tRNA synthetase. Biochem J. 1980;191(1):209–219. doi: 10.1042/bj1910209. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 238.Brown MJ, Mensah LM, Doyle ML, Broom NJ, Osbourne N, Forrest AK, Richardson CM, O’Hanlon PJ, Pope AJ. Rational design of femtomolar inhibitors of isoleucyl tRNA synthetase from a binding model for pseudomonic acid-A. Biochemistry. 2000;39(20):6003–6011. doi: 10.1021/bi000148v. [DOI] [PubMed] [Google Scholar]
- 239.Yanagisawa T, Lee JT, Wu HC, Kawakami M. Relationship of protein structure of isoleucyl-tRNA synthetase with pseudomonic acid resistance of Escherichia coli. A proposed mode of action of pseudomonic acid as an inhibitor of isoleucyl-tRNA synthetase. J Biol Chem. 1994;269(39):24304–24309. [PubMed] [Google Scholar]
- 240.Rock FL, Mao W, Yaremchuk A, Tukalo M, Crepin T, Zhou H, Zhang YK, Hernandez V, Akama T, Baker SJ, et al. An antifungal agent inhibits an aminoacyl-tRNA synthetase by trapping tRNA in the editing site. Science. 2007;316(5832):1759–1761. doi: 10.1126/science.1142189. [DOI] [PubMed] [Google Scholar]
- 241.Baker SJ, Zhang YK, Akama T, Lau A, Zhou H, Hernandez V, Mao W, Alley MR, Sanders V, Plattner JJ. Discovery of a new boron-containing antifungal agent, 5-fluoro-1,3-dihydro-1-hydroxy-2,1-benzoxaborole (AN2690), for the potential treatment of onychomycosis. J Med Chem. 2006;49(15):4447–4450. doi: 10.1021/jm0603724. [DOI] [PubMed] [Google Scholar]
- 242.Greenberg Y, King M, Kiosses WB, Ewalt K, Yang X, Schimmel P, Reader JS, Tzima E. The novel fragment of tyrosyl tRNA synthetase, mini-TyrRS, is secreted to induce an angiogenic response in endothelial cells. FASEB J. 2008;22(5):1597–1605. doi: 10.1096/fj.07-9973com. [DOI] [PubMed] [Google Scholar]
- 243.Wakasugi K, Slike BM, Hood J, Otani A, Ewalt KL, Friedlander M, Cheresh DA, Schimmel P. A human aminoacyl-tRNA synthetase as a regulator of angiogenesis. Proc Natl Acad Sci U S A. 2002;99(1):173–177. doi: 10.1073/pnas.012602099. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 244.Otani A, Slike BM, Dorrell MI, Hood J, Kinder K, Ewalt KL, Cheresh D, Schimmel P, Friedlander M. A fragment of human TrpRS as a potent antagonist of ocular angiogenesis. Proc Natl Acad Sci U S A. 2002;99(1):178–183. doi: 10.1073/pnas.012601899. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 245.Zhou Q, Kapoor M, Guo M, Belani R, Xu X, Kiosses WB, Hanan M, Park C, Armour E, Do MH, et al. Orthogonal use of a human tRNA synthetase active site to achieve multifunctionality. Nature Struct Mol Biol. 2010;17(1):57–61. doi: 10.1038/nsmb.1706. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 246.Dorrell MI, Aguilar E, Scheppke L, Barnett FH, Friedlander M. Combination angiostatic therapy completely inhibits ocular and tumor angiogenesis. Proc Natl Acad Sci U S A. 2007;104(3):967–972. doi: 10.1073/pnas.0607542104. [DOI] [PMC free article] [PubMed] [Google Scholar]
- 247.Arnez JG, Dock-Bregeon AC, Moras D. Glycyl-tRNA synthetase uses a negatively charged pit for specific recognition and activation of glycine. J Mol Biol. 1999;286(5):1449–1459. doi: 10.1006/jmbi.1999.2562. [DOI] [PubMed] [Google Scholar]
- 248.Mursinna RS, Martinis SA. Rational design to block amino acid editing of a tRNA synthetase. J. Am. Chem. Soc. 2002;124(25):7286–7287. doi: 10.1021/ja025879s. [DOI] [PubMed] [Google Scholar]








